Deep eutectic solvent
Deep eutectic solvents (DESs) are fluid mixtures typically formed by combining two or three inexpensive components, often safe for common formulations—such as a hydrogen bond acceptor (e.g., a quaternary ammonium salt like choline chloride) and a hydrogen bond donor (e.g., urea, glycerol, or a metal halide)—through intermolecular interactions, primarily hydrogen bonding, that result in a eutectic mixture with a significantly depressed melting point lower than that of any individual component.[1][2] These solvents remain liquid at temperatures below 100 °C, often at or near room temperature, and were first reported in 2001 by Abbott and colleagues using choline chloride and urea.[2][3] DESs exhibit physicochemical properties akin to those of room-temperature ionic liquids but are distinguished by their simpler preparation from readily available precursors that are often biodegradable and low-toxicity, especially in bio-based variants, making them a greener alternative for various applications.[1] Key characteristics include low volatility, high thermal stability, tunable polarity and viscosity (which can be adjusted by adding water or varying component ratios), wide liquidus temperature ranges, and high solubility for diverse solutes such as metal salts, organic compounds, and biomolecules.[2][3] They generally possess high densities, moderate conductivity, and non-flammable natures, though their elevated viscosities compared to conventional solvents can sometimes limit mass transfer in processes.[1][4] Classified into types based on composition, DESs include Type I (quaternary ammonium salts with metal halides), Type II (with hydrated metal halides), Type III (with hydrogen bond donors like amides or alcohols, often natural deep eutectic solvents or NADES), and Type IV (metal halides with urea derivatives); hydrophobic variants and therapeutic DESs (THEDES) have also emerged for specialized uses.[2][3] Their applications span metal electrodeposition and processing, organic synthesis, extraction of bioactive compounds (e.g., phenolics and flavonoids from natural sources), carbon dioxide capture, biotransformations, and pharmaceutical formulations for enhanced drug solubility and delivery.[1][2][3] Ongoing research emphasizes their eco-friendly profile and versatility in sustainable chemistry, with as of 2025 advancements including responsive DES and expanded uses in biomass pretreatment and food safety sensors.[4][5]Definition and Fundamentals
Definition
Deep eutectic solvents (DESs) are eutectic mixtures formed from Lewis or Brønsted acids and bases that contain a hydrogen bond network, resulting in significant depression of the freezing point compared to the individual components.[2] These mixtures typically involve a quaternary ammonium salt, such as choline chloride, complexed with a hydrogen bond donor like urea or a metal salt, where the interactions delocalize charges and stabilize the liquid state at lower temperatures.[2] A classic example is the 1:2 molar ratio mixture of choline chloride (decomposition point 302°C) and urea (melting point 133°C), which forms a clear liquid with a melting point of approximately 12–25°C, depending on water content (e.g., 12°C with ~0.65 wt% water, 23–25°C for low-water conditions) due to the extensive hydrogen bonding between the chloride anion and urea molecules.[6][2][7] Unlike ionic liquids, which are pure salts composed of discrete ions, DESs are multicomponent mixtures that exhibit similar physical properties but are easier and cheaper to prepare without the need for purification.[2] In contrast to volatile organic solvents, DESs possess low vapor pressure and are non-flammable, enhancing their safety in applications.[2][8] General advantages of DESs include their tunability through component selection, low toxicity, and biodegradability, making them attractive "green" alternatives in chemical processes.[2]Eutectic Behavior
Eutectic mixtures consist of two or more components that, upon mixing, exhibit a eutectic point in their phase diagram where the liquidus and solidus curves intersect, resulting in a sharp melting temperature lower than that of either pure component. This depression arises from intermolecular interactions that stabilize the liquid phase relative to the crystalline solids, preventing solidification until a lower temperature is reached. In the context of deep eutectic solvents (DES), these interactions lead to a pronounced eutectic behavior, forming a stable liquid over a wide temperature range, often at or near room temperature.[9] The primary mechanism for freezing point depression in DES involves hydrogen bonding between a hydrogen bond acceptor (typically a quaternary ammonium salt) and a hydrogen bond donor (such as a carboxylic acid or urea), which disrupts the ordered structure of the individual components. In ionic components like choline chloride, hydrogen bonding with the donor delocalizes the charge on halide anions, effectively reducing the lattice energy of the ionic solid and lowering the energy barrier for melting. This lattice disruption is particularly evident in Type I and Type III DES, where coordination between the donor and ionic species prevents recrystallization, enhancing the depth of the eutectic.[10] For ideal eutectic mixtures, the melting point depression can be approximated using colligative property principles derived from the equality of chemical potentials between the solid and liquid phases. Consider a binary mixture where component 1 (the solvent) has a pure melting temperature T_m, enthalpy of fusion \Delta H_f, and mole fraction x_1 in the liquid. At the depressed melting temperature T = T_m - \Delta T, the chemical potential of pure solid component 1 equals that in the ideal solution: \mu_1^\text{solid}(T) = \mu_1^\text{liquid}(T). The solid chemical potential is \mu_1^\text{solid}(T) = \mu_1^\text{liquid}(T_m) - \int_T^{T_m} \frac{\Delta H_f}{T'} dT', assuming constant \Delta H_f. Approximating the integral as \Delta H_f \ln(T_m / T) and using the liquid chemical potential \mu_1^\text{liquid}(T) = \mu_1^\text{liquid}(T_m) + RT \ln x_1, the equation simplifies to \Delta H_f \ln(T / T_m) = RT \ln x_1. For small \Delta T, \ln(T / T_m) \approx -\Delta T / T_m, yielding \Delta T = -\frac{R T_m^2}{\Delta H_f} \ln x_1. Since x_1 = 1 - x_2 and for small solute mole fraction x_2, \ln x_1 \approx -x_2, the ideal depression becomes \Delta T \approx \frac{R T_m^2}{\Delta H_f} x_2, where R is the gas constant.[9] This equation illustrates the scale of depression in DES; for example, in a choline chloride-urea mixture at a 1:2 molar ratio (x_2 \approx 0.67), the observed \Delta T exceeds 100°C compared to the components' melting points, though non-ideal interactions amplify the effect beyond the ideal prediction. Factors influencing the eutectic depth include the optimal component ratio, which determines the mole fraction at the eutectic point and maximizes stabilization, as well as the nature of interactions—stronger hydrogen bonding or ionic delocalization yields deeper depressions than weaker van der Waals forces, while molecular donors promote more pronounced effects in ionic systems than in purely molecular ones.[10][9]Classification and Types
Conventional DES
Conventional deep eutectic solvents (DES) are synthetic mixtures classified into four main types based on their component compositions, primarily involving quaternary ammonium salts, metal chlorides, or hydrogen bond donors (HBDs).[2] This classification, introduced by Abbott and colleagues, highlights their structural diversity and tunability for various synthetic applications.[2] Type I DES consist of a quaternary ammonium salt, such as choline chloride (ChCl), combined with a metal chloride like ZnCl₂ or AlCl₃.[2] A representative example is ChCl:ZnCl₂ at a 1:2 molar ratio, which forms a stable liquid at room temperature suitable for Lewis acid-catalyzed syntheses, such as Diels-Alder reactions, though these mixtures can be sensitive to hydrolysis due to the anhydrous metal salts.[2] Type II DES incorporate a quaternary ammonium salt with a hydrated metal chloride, exemplified by ChCl:CrCl₃·6H₂O at a 1:2 molar ratio (choline chloride to the hexahydrate). These are notably air- and moisture-stable, enabling their use in metal electrodeposition and esterification reactions without degradation.[2] Type III DES are metal-free and formed by a quaternary ammonium salt paired with an HBD, such as urea or glycerol; common examples include ChCl:urea (1:2 molar ratio) and ChCl:glycerol (1:2 molar ratio), both exhibiting good thermal stability and low toxicity for applications in N-alkylation and biodiesel-related syntheses.[2] Type IV DES involve a metal chloride hydrate mixed with an HBD, without quaternary ammonium salts; a typical formulation is ZnCl₂:urea at a 1:3.5 molar ratio, which provides moderate stability and is employed in processes like aluminum plating and certain carbonylation reactions.[11] A more recent addition to the classification is Type V DES, which are formed by non-ionic hydrogen bond acceptors (HBAs) and donors (HBDs), such as monoterpenoids (e.g., menthol:thymol at 1:1 molar ratio) or fatty acids with terpenes (e.g., decanoic acid:menthol at 1:1), without halides or metals. These are often hydrophobic, enabling applications in non-aqueous environments like oil-water separations, and represent a greener, chloride-free alternative.[12][13] Compared to traditional ionic liquids, conventional DES offer advantages in cost-effectiveness, as their components are inexpensive and commercially available, and ease of tuning physicochemical properties through simple variation of molar ratios or component selection.[2]Natural Deep Eutectic Solvents
Natural deep eutectic solvents (NADES) are eutectic mixtures composed of two or more natural metabolites that form a viscous liquid phase at temperatures below the melting points of their individual components, serving as a sustainable alternative to conventional solvents.[14] These solvents typically include bio-based compounds such as sugars (e.g., glucose, fructose, sucrose), organic acids (e.g., citric acid, malic acid, lactic acid), amino acids (e.g., proline), and choline derivatives, often combined with water as a third component to achieve the eutectic mixture.[14][15] NADES were first conceptualized in 2011 by Choi et al. as a means to solubilize plant metabolites that exhibit poor water solubility, such as rutin and paclitaxel, thereby addressing gaps in understanding cellular metabolism and physiology.[14] Representative examples include glucose:choline chloride:water in a 1:1:1 molar ratio and citric acid:choline chloride in 1:2 or 1:3 ratios, which demonstrate the formation of stable liquids through hydrogen bonding interactions.[14] Another example is betaine:sorbitol:water (1:1:3 molar ratio), a ternary mixture that highlights the versatility of quaternary ammonium salts like betaine as hydrogen bond acceptors in NADES formulations.[16] Water plays a crucial role in NADES by being strongly retained within the mixture (often up to 6-30% by weight), enhancing thermodynamic stability and mimicking the hydrated intracellular fluids found in living organisms, where it facilitates enzyme activity and metabolite dissolution without evaporation under vacuum.[14] This hydration aspect positions NADES as analogs to cellular compartments, potentially representing a third liquid phase alongside water and lipids in biological systems.[14] Compared to conventional deep eutectic solvents, NADES offer superior environmental benefits due to their derivation from renewable natural sources, resulting in low toxicity profiles and high biodegradability, which minimize ecological impacts in green chemistry applications.[15][17] These attributes make NADES particularly advantageous for sustainable processes, as they degrade readily in the environment without persistent residues.[15]Hydrophobic and Therapeutic DES
Hydrophobic deep eutectic solvents (HDES) are a specialized class, often based on Type V formulations using long-chain fatty acids or terpenes (e.g., lauric acid:decylamine 1:1), which are immiscible with water and useful for extracting non-polar compounds or in biphasic systems.[13] Therapeutic deep eutectic solvents (THEDES) incorporate active pharmaceutical ingredients (APIs) with safe co-formers (typically Type III-like), enhancing drug solubility and delivery while maintaining biocompatibility.[18]History
Discovery and Early Development
The discovery of deep eutectic solvents (DES) traces back to 2001, when Andrew P. Abbott and colleagues at the University of Leicester reported the formation of low-melting eutectic mixtures by combining quaternary ammonium salts, such as choline chloride, with zinc chloride (ZnCl₂). These mixtures displayed significantly depressed freezing points—down to 23–25 °C for certain compositions—compared to the high melting points of the individual components (choline chloride at 302 °C and ZnCl₂ at 290 °C), enabling liquid states at or near room temperature. This work laid the groundwork for DES as a class of ionic liquid analogs, highlighting their potential as moisture-stable, Lewis-acidic media for chemical processes.[19] In 2003, Abbott's group expanded on this by demonstrating that a 1:2 molar mixture of choline chloride and urea formed a clear, stable DES at ambient temperatures, with a melting point of 12 °C. Dubbed "Reline," this solvent exhibited novel properties, including the ability to dissolve metal oxides like ZnO and CuO at elevated temperatures, making it suitable for metal processing applications such as electropolishing and extraction. Early experiments focused on its use in electrochemical environments, where it supported the dissolution of metals and facilitated processes like the electrodeposition of zinc from ZnCl₂-based systems.[6] From 2003 to 2010, research positioned DES as economical alternatives to traditional ionic liquids, which originated with Paul Walden's 1914 synthesis of ethylammonium nitrate as the first room-temperature ionic liquid. DES offered advantages in cost and simplicity, using inexpensive, biorenewable components without requiring purification steps typical of ionic liquids. Initial applications emphasized electrochemistry and extraction; for instance, Abbott et al. in 2004 explored DES formed with carboxylic acids for metal oxide solubility, enabling selective extraction and potential electrodeposition scenarios. By the mid-2000s, studies demonstrated DES utility in aluminum-related electrochemistry, achieving deposits with high purity and efficiency in choline chloride-based media. However, early investigations consistently noted challenges, including high viscosities (often exceeding 100 mPa·s at room temperature), which limited mass transport and practical scalability in electrochemical setups.[2][20][21]Key Milestones and Evolution
Building upon the early foundational research by Abbott and colleagues on choline chloride-based eutectic mixtures, the field of deep eutectic solvents (DES) saw significant diversification in the 2010s through the introduction of bio-derived and environmentally benign variants.[2] A pivotal advancement occurred in 2011 with the proposal of natural deep eutectic solvents (NADES), introduced by Choi et al. as mixtures of natural metabolites such as sugars, amino acids, and organic acids that form eutectic liquids at ambient temperatures, enabling efficient extraction of natural products like flavonoids and alkaloids from plant matrices without toxic residues.[14] This innovation shifted focus toward biocompatible solvents, expanding DES applicability in extraction processes for bioactive compounds.[22] The year 2014 marked a surge in research interest following the comprehensive review by Smith et al., which systematically outlined DES formation mechanisms, physicochemical properties, and diverse applications ranging from metal electrodeposition to organic synthesis, thereby establishing DES as viable ionic liquid alternatives and catalyzing over a decade of subsequent studies.[2] In the mid-2010s, research expanded to hydrophobic DES formulations, exemplified by Florindo et al.'s 2017 development of stable, water-immiscible systems using decanoic acid as the hydrogen bond donor paired with quaternary ammonium salts like tetrabutylammonium bromide, which demonstrated low water solubility (~0.3 wt%) and enabled extraction of pollutants such as neonicotinoid pesticides from aqueous streams.[23] Concurrently, Type IV DES—comprising metal chloride hydrates (e.g., ZnCl₂·3H₂O) with hydrogen bond donors like urea—gained traction for their tunable viscosity and conductivity in electrochemical applications, while efforts toward metal-free variants emerged to enhance biocompatibility, such as carbohydrate-based eutectics reported around 2019 that avoided metallic components entirely.[2] From 2019 to 2022, DES research emphasized sustainability, particularly in biomass processing, where choline chloride-oxalic acid mixtures achieved up to 58% delignification of lignocellulosic feedstocks like wheat straw under mild conditions, facilitating enzymatic hydrolysis for biofuel production.[24] This period also featured key EU-funded initiatives under Horizon 2020, such as the NADES4CP project (2017–2021), which developed temperature- and halide-enhanced DES for scalable chemical engineering processes, including CO₂ capture and biocatalysis, underscoring DES as green alternatives to volatile organic compounds. Post-2022, advancements included the development of responsive deep eutectic solvents (RDES) with switchable properties for controlled separations and catalysis, as highlighted in 2024 reviews, further integrating DES into circular economy applications.[5]Preparation
Synthesis Methods
Deep eutectic solvents (DESs) are typically prepared by combining a hydrogen bond acceptor, such as a quaternary ammonium salt like choline chloride, with a hydrogen bond donor, such as urea or a carboxylic acid, in specific molar ratios to form a homogeneous liquid phase.[2] The most common synthesis method is the heating approach, where the solid components are mixed and heated to temperatures between 50°C and 100°C under constant stirring until a clear, stable liquid forms, often within 1-2 hours depending on the mixture. This method relies on the formation of hydrogen bonds that depress the melting point below that of the individual components, avoiding thermal degradation by keeping temperatures below the stability limits of the constituents. For instance, the prototypical DES known as "reline" is synthesized by mixing choline chloride and urea in a 1:2 molar ratio at around 80°C, resulting in a colorless liquid with a melting point of 12°C.[25][26] Mechanochemical methods provide non-thermal alternatives, particularly suitable for heat-sensitive components. These include grinding the solid components together at room temperature until a homogeneous liquid forms, or using continuous processes like twin-screw extrusion for scalable production. For example, reline can be prepared by grinding choline chloride and urea (1:2) without heating, achieving high-purity DES in minutes. Twin-screw extrusion enables efficient, solvent-free synthesis in multi-kg/h quantities, with precise control over parameters like screw speed and temperature to maintain eutectic composition.[26][27][28] Non-heating methods are employed for heat-sensitive components, particularly in natural deep eutectic solvents (NADES), to prevent degradation. In vacuum evaporation, the components are first dissolved in a volatile solvent like water to form an aqueous mixture, which is then subjected to reduced pressure and mild heating (below 50°C) to remove the solvent, yielding the DES as a residue. This technique is suitable for hydrophilic DESs and ensures high purity by minimizing exposure to elevated temperatures. Alternatively, freeze-drying involves preparing an aqueous solution of the components, freezing it at temperatures as low as -77°C, and then sublimating the ice under vacuum to obtain the DES directly, which is particularly effective for NADES containing sugars or amino acids.[25][29][30] Scalability of DES synthesis can be achieved through batch or continuous processes, with batch methods being simpler for laboratory-scale production using standard glassware, while continuous flow systems enhance efficiency for industrial applications by allowing precise control over mixing and temperature. In batch processes, small volumes (e.g., 100 mL to liters) are prepared in stirred reactors, but scaling up requires attention to heat transfer and uniformity to maintain the eutectic composition. Continuous synthesis, such as in tubular reactors or twin-screw extruders, has been demonstrated for high-purity DESs like choline chloride-based mixtures, achieving production rates up to multi-kg/h with reduced energy input and consistent quality. Purity control is critical in both approaches, as impurities can shift the eutectic point and alter properties; thus, high-grade reagents and post-synthesis filtration are recommended to achieve water content below 0.5 wt% and avoid phase separation.[27][28] Safety considerations during DES preparation emphasize the low volatility and non-flammability of most formulations, reducing risks compared to traditional organic solvents, though handling quaternary salts requires gloves and ventilation to avoid skin irritation. Equipment typically includes magnetic stirrers or overhead mixers for agitation, oil baths or hot plates for heating under inert atmospheres (e.g., nitrogen) to protect air-sensitive components, and rotary evaporators or freeze-dryers for non-thermal methods, ensuring operations in well-ventilated fume hoods.[2][31][32]Characterization Techniques
Characterization of deep eutectic solvents (DES) involves a suite of analytical techniques to confirm their formation through eutectic depression, verify intermolecular interactions, and evaluate key physicochemical properties such as thermal stability and transport behavior. These methods are crucial for distinguishing true DES from simple mixtures and ensuring their suitability for applications. Thermal, spectroscopic, electrochemical, and rheological analyses are standard, often applied post-synthesis to assess purity and homogeneity.[33] Thermal analysis employs differential scanning calorimetry (DSC) and thermogravimetric analysis (TGA) to probe phase transitions and stability. DSC measures the melting point and glass transition temperature, revealing significant depressions (often 50–200 K) from ideal eutectic behavior, which confirms the non-ideal interactions defining DES formation; for example, in choline chloride-urea systems, DSC identifies the eutectic melting point at approximately 12 °C.[34][33] TGA assesses thermal decomposition by monitoring mass loss, typically showing onset temperatures of 135–250°C for common DES like those based on choline chloride and organic acids, indicating good short-term stability under heating.[35][33] These techniques together establish the temperature range for DES usability without phase separation or degradation.[34] Spectroscopic methods elucidate molecular-level interactions, particularly hydrogen bonding between hydrogen bond donors (HBD) and acceptors (HBA). Fourier-transform infrared (FTIR) spectroscopy detects shifts in vibrational bands, such as broadening or shifting of O-H (around 3300 cm⁻¹) and C=O (1700–1750 cm⁻¹) stretches, confirming HBD-HBA complexation; in tetrabutylammonium bromide-nonanoic acid mixtures, FTIR verifies bonding at optimal ratios.[34][35] Raman spectroscopy complements FTIR by highlighting symmetric vibrations less affected by dipole changes, revealing ion pairing and lattice disruption in DES like choline chloride-glycerol.[33] Nuclear magnetic resonance (NMR) spectroscopy, including ¹H and ¹³C variants, probes component dynamics and interactions through chemical shift perturbations and peak broadening, demonstrating restricted mobility in the liquid state; for instance, in menthol-based DES, NMR confirms fatty acid integration via upfield shifts.[35][33] Electrochemical techniques focus on conductivity to characterize ionic mobility. Electrochemical impedance spectroscopy (EIS) measures specific conductivity by applying an AC signal and analyzing the Nyquist plot, yielding values typically 0.1–10 mS/cm at 25°C for DES, influenced by viscosity and ion dissociation; this verifies the solvent's electrolytic potential without electrodeposition issues.[33] Lower conductivities in viscous DES arise from strong H-bond networks, as seen in choline chloride-ethylene glycol systems.[35] Rheological analysis uses rotational viscometers to determine viscosity profiles under shear, essential for understanding flow and processing. DES exhibit high viscosities (10–5000 mPa·s at 25°C), often Newtonian but sometimes shear-thinning, decreasing exponentially with temperature (e.g., activation energies of 30–60 kJ/mol); in TBAB-nonanoic acid DES, viscosity drops from 124 mPa·s at 1:1 ratio to 5 mPa·s at higher HBD content at 40°C, reflecting weakened interactions.[34][33][35] These measurements guide dilution strategies to optimize handling.Properties
Physical Properties
Deep eutectic solvents (DES) exhibit densities typically ranging from 1.0 to 1.5 g/cm³ at ambient temperatures, with values for common choline chloride (ChCl)-based DES falling between 1.12 and 1.24 g/cm³ at 298 K.[2] These densities decrease with increasing temperature and can be influenced by water content, where small additions may slightly increase density due to hydrogen bonding interactions.[36] Viscosity in DES is notably high, often spanning 10 to 1000 cP or more at room temperature, exceeding that of many ionic liquids; for instance, ChCl:ethylene glycol (1:2) has a viscosity of 36 cP, while ChCl:urea (1:2) reaches 632 cP at 298 K.[2] This property is highly sensitive to temperature, decreasing exponentially per the Arrhenius model, and to water content, which can reduce viscosity by disrupting the hydrogen bond network.[2][36] Ionic conductivity of DES generally lies in the range of 0.1 to 10 mS/cm at 298 K, lower than conventional ionic liquids owing to the elevated viscosity that impedes ion mobility; examples include 0.75 mS/cm for ChCl:urea (1:2) and 7.61 mS/cm for ChCl:ethylene glycol (1:2).[2] Conductivity increases with temperature as viscosity drops. DES possess near-negligible vapor pressure, often below 1 mmHg (or ~0.13 kPa), contributing to their low volatility and environmental stability; for instance, certain DES mixtures show pressures of 2–60 Pa (0.015–0.45 mmHg) at 343–393 K.[2][37] Surface tension values for DES are typically 40–60 mN/m at ambient conditions, though some reach up to 77 mN/m depending on composition, decreasing with temperature; ChCl:ethylene glycol (1:2) measures 49 mN/m at 298 K.[2] Thermal stability is robust, with many DES exhibiting decomposition onset temperatures above 200°C as determined by thermogravimetric analysis; ChCl-based DES with hydrogen bond donors like urea or glycerol often surpass this threshold.[38] The physical properties of DES are highly tunable through adjustments in the hydrogen bond acceptor (HBA) to hydrogen bond donor (HBD) molar ratios, allowing optimization for specific applications. For example, in ChCl:D-fructose, viscosity drops from 14,347 cP (1:1) to lower values with higher ratios. The following table summarizes properties for select common DES at 298 K:| DES Composition (HBA:HBD) | Density (g/cm³) | Viscosity (cP) | Conductivity (mS/cm) | Source |
|---|---|---|---|---|
| ChCl:Urea (1:2) | 1.24 | 632 | 0.75 | https://pubs.acs.org/doi/10.1021/cr300162p |
| ChCl:Ethylene Glycol (1:2) | 1.12 | 36 | 7.61 | https://pubs.acs.org/doi/10.1021/cr300162p |
| ChCl:Glycerol (1:2) | 1.19 | 149 | 1.05 | https://pubs.acs.org/doi/10.1021/cr300162p |
| ChCl:Citric Acid (1:1) | 1.23 | 10,224 | - | https://pubs.acs.org/doi/10.1021/acs.jced.3c00706 |