Fact-checked by Grok 2 weeks ago

Solvation

Solvation is the process by which molecules surround and interact with solute particles—ions or molecules—through intermolecular forces, thereby stabilizing the solute within the . This interaction, often involving the formation of a around the solute, is essential for the of substances and occurs when the solute-solvent attractions overcome the solute-solute and -solvent attractions. Solvation describes the molecular-level interactions, distinct from , which is the maximum amount of solute that can dissolve in a at . When the is , the process is specifically termed , where molecules orient their polar ends toward charged or polar solute . The solvation process can be broken down into three main steps: the endothermic separation of solvent molecules from each other, the endothermic separation of solute particles, and the exothermic formation of solute-solvent interactions. The overall change (ΔH_soln) depends on the balance of these energies; if the exothermic step dominates, solvation is exothermic (e.g., dissolution of in , releasing heat), whereas if endothermic steps prevail, it is endothermic (e.g., dissolution of in , absorbing heat). changes, particularly the increased disorder from dispersing solute particles, often drive the spontaneity of solvation even when it is endothermic. Solvation plays a pivotal role in chemistry and biochemistry, influencing , reaction rates, and molecular behavior in solutions. It underlies phenomena such as the of ions in aqueous environments, the folding of proteins through hydrophobic and hydrophilic interactions, and the of solutions. Understanding solvation is crucial for applications in fields like , pharmaceuticals, and , where solvent-solute interactions determine the efficacy of processes like or pollutant transport.

Introduction

Definition and Process

Solvation is the process by which molecules or ions of a solute interact with and become surrounded by molecules of a , forming a cluster known as the that stabilizes the solute through intermolecular forces. This interaction typically involves weak bonds, such as electrostatic attractions, allowing the solute to disperse within the and form a homogeneous . The extent of solvation depends on the chemical nature of both solute and , influencing the and reactivity of the system. At the molecular level, solvation begins with the reorganization of molecules around the solute, where dipoles or polar groups orient toward the solute to minimize energy. The primary , or first , consists of molecules in direct, intimate contact with the solute, often forming coordination bonds in the case of metal ions or bonds with polar solutes. Beyond this, secondary solvation shells form with more distant molecules influenced by the primary layer, held together primarily by forces and weaker electrostatic interactions. These layered structures dynamically adjust to the solute's charge and size, enhancing through collective solvent effects. A prominent example of solvation is , the specific case where acts as the , with molecules forming bonds and orienting their to solvate polar or charged solutes. In ionic solutions, such as aqueous NaCl, cations like Na⁺ attract the negative oxygen ends of molecules in the primary shell, while anions like Cl⁻ interact with the positive ends, creating oriented layers that screen the ions' charges and facilitate . This process is essential for the behavior of electrolytes in solution. The foundational recognition of solvation's role in electrolyte solutions dates to Svante Arrhenius's 1887 theory of electrolytic dissociation, where he described how ions in dilute aqueous solutions are hydrated, linking this hydration to observed electrical conductivity and solution properties.

Distinction from Solubility

Solvation and solubility are related but distinct concepts in solution chemistry, with solvation describing the molecular-level interactions that stabilize solute particles through solvent association, while solubility quantifies the equilibrium state of a saturated solution. Solvation describes the stabilizing interactions between solute and solvent, which are part of the dissolution process. The kinetics of dissolution, such as the rate at which solute dissolves (e.g., in mol/s), involve solvation but are distinct from the equilibrium property of solubility, defined as the maximum amount of solute that can coexist with the undissolved phase in a solution, measured in concentration units such as mol/L or g/100 mL. This distinction highlights that solvation focuses on the mechanism of solute-solvent interactions, whereas solubility reflects the thermodynamic limit of dissolution under equilibrium conditions. According to IUPAC definitions, solvation encompasses any stabilizing between a solute (or solute moiety) and the , including similar interactions with groups on insoluble materials, emphasizing the role of intermolecular forces in solute stabilization. , however, is the analytical composition of a saturated , expressed as the proportion of the designated solute in the , without to the underlying . These definitions underscore that solvation can occur independently of achieving high ; for instance, transient solvation complexes may form briefly without leading to a measurable increase in dissolved solute concentration, as seen in interactions with surface groups on insoluble substrates. Conversely, inherently requires effective solvation to stabilize dissolved but is also governed by additional factors, such as the energy needed to overcome the solute's structure in solids. A key molecular feature of solvation is the , which represents the average number of solvent molecules directly bound to the solute in the first ; for example, the sodium (Na⁺) in typically exhibits a coordination number of approximately 5.5 to 6 oxygen atoms from water molecules. This shell forms rapidly through electrostatic interactions, illustrating solvation's dynamic nature. An illustrative case is the solvation of gaseous ions upon introduction into a solvent: isolated ions from the gas phase solvate almost instantaneously as solvent molecules cluster around them, driven by ion-dipole forces, yet the overall of the corresponding solid salt may remain low if the crystal exceeds the solvation energy gained. In such scenarios, the initial solvation step occurs efficiently, but the equilibrium is limited by the balance between lattice disruption and solvation stabilization.

Intermolecular Interactions

Types of Solvent-Solute Forces

The primary intermolecular forces responsible for solvation include ion-dipole interactions, which dominate when ionic solutes are present in polar s, dipole-dipole interactions and hydrogen bonding for polar solutes, and London dispersion forces for nonpolar solutes. These forces facilitate the organization of solvent molecules around the solute, forming a that stabilizes the dissolved species through electrostatic and van der Waals attractions. For ionic solutes, ion-dipole interactions arise from the attraction between the charged and the partial charges on polar molecules, leading to the orientation of dipoles with their positive ends toward anions and negative ends toward cations. This alignment creates a structured first , where the of the is given by U = -\frac{q \mu \cos\theta}{4\pi\epsilon_0 r^2} where q is the charge, \mu is the of the , \theta is the angle between the dipole and the line connecting the to the dipole center, \epsilon_0 is the , and r is the distance between the and the dipole center. In protic s like , polar solutes engage in dipole-dipole s supplemented by hydrogen bonding, where molecules form directional networks that bridge the solute's polar groups, enhancing shell stability through cooperative effects. Nonpolar solutes, lacking permanent dipoles, rely on London forces, which are induced temporary dipoles arising from correlated electron fluctuations, allowing weak but cumulative attractions in the . In certain systems, charge transfer, where the direction depends on the type—electron donation from to cations or from anions to —can contribute to -solute interactions, particularly in solvated ionic clusters, altering the effective charge distribution and strengthening binding. Solvatochromism provides experimental evidence for these varying interactions, as shifts in the electronic spectra of solutes reflect changes in the local environment, such as or hydrogen-bonding capacity. For instance, around hydrophobic groups in , molecules adopt a tetrahedral hydrogen-bonding arrangement that maintains bulk-like ordering while excluding the solute, minimizing disruption to the network. In low-dielectric s, pairing occurs as a consequence of weakened solvation, where oppositely charged ions associate closely, reducing the separation of the solvation shells.

Solvent Properties and Classification

Solvents are broadly classified based on their ability to participate in hydrogen bonding and their overall polarity, which directly influence their capacity to solvate different types of solutes. Protic solvents contain labile protons attached to electronegative atoms, such as oxygen or nitrogen, enabling them to act as hydrogen bond donors; examples include water and alcohols like methanol. In contrast, aprotic solvents lack such protons and cannot donate hydrogen bonds, though they may accept them; representative aprotic solvents are acetone and dimethyl sulfoxide (DMSO). This distinction affects solvation efficiency, as protic solvents stabilize charged or polar solutes through hydrogen bonding, while aprotic solvents are better suited for non-hydrogen-bonding interactions. Solvents are further categorized by polarity into polar and nonpolar types, with polar solvents exhibiting significant moments or charge separation that enhances their ability to dissolve ionic or polar solutes. Nonpolar solvents, such as , have minimal moments and preferentially solvate nonpolar molecules via weak forces. exemplifies amphiprotic behavior within the protic category, as it can both donate and accept bonds due to its hydroxyl group, allowing versatile solvation of a range of solutes. Key properties quantifying solvent behavior include the dielectric constant (ε), which measures a solvent's ability to screen electrostatic interactions; for at 25°C, ε = 78.5, indicating strong and effective solvation of ions. The Gutmann donor number (DN) assesses a solvent's basicity toward cations, with values derived from calorimetric measurements of formation with SbCl₅; higher DN values, like water's DN = 18, signify stronger cation solvation. Complementarily, the acceptor number (AN) evaluates electrophilic properties via ³¹P NMR shifts with triethylphosphine oxide, where higher AN (e.g., water's AN = 54.8) reflects better anion solvation. The Kamlet-Taft parameters provide a multiparametric : π* for dipolarity/, α for donation, and β for acceptance; these enable prediction of solvatochromic shifts and solute selectivity in diverse solvents. The index (P'), based on interactions with probe solutes, influences solute selectivity by ranking from nonpolar (e.g., , P' = 0.0) to highly polar (e.g., , P' = 10.2), guiding choices for specific solvation tasks. Non-aqueous like ionic liquids offer specialized solvation due to their tunable and low ; for instance, protic ionic liquids exhibit variable Kamlet-Taft parameters that allow selective of both polar and nonpolar compounds. Supercritical CO₂, with its low dielectric constant (ε ≈ 1.6 near critical point), serves as a nonpolar for hydrophobic solutes, enhanced by adjustable for processes. Solvent and impact the dynamics of solvation by governing solute into solvation shells; higher , as in ionic liquids (often >10 ), slows rates, potentially limiting solvation , while variations modulate local solvent structuring around solutes. In mixtures, even small changes can significantly alter mobility without disrupting core solvation shells, affecting overall solvation efficiency.
Solvent ClassExamplesKey Property ExampleInfluence on Solvation
Polar ProticWater, Methanolε = 78.5 (water); α > 0Strong ion stabilization via H-bonding
Polar AproticAcetone, DMSOβ ≈ 0.5–0.8; π* ≈ 0.6–1.0Enhanced nucleophile reactivity
NonpolarHexane, Supercritical CO₂ε < 5; P' ≈ 0Preferential nonpolar solute dissolution
Ionic Liquids[Emim][BF₄]Tunable DN/ANSelective for mixed polarity solutes

Thermodynamic Aspects

Solvation Energy

Solvation energy quantifies the energetic cost or benefit associated with transferring a solute from the gas phase to a , primarily arising from three key components: electrostatic stabilization due to solute-solvent charge interactions, the positive cost of forming a in the solvent to accommodate the solute, and attractive forces between the solute and solvent molecules. The electrostatic term dominates for charged solutes, providing stabilization through polarization of the solvent medium, while the cavity formation term opposes solvation by disrupting solvent-solvent interactions, and contributes a smaller attractive component via van der Waals forces. The model provides a foundational description of the electrostatic contribution to solvation energy, treating the solute as a charged of r embedded in a with \epsilon. This model calculates the change for charging the in the relative to as: \Delta G_{\text{Born}} = -\frac{N_A z^2 e^2}{8\pi \epsilon_0 r} \left(1 - \frac{1}{\epsilon}\right) where N_A is Avogadro's number, z is the charge, e is the , and \epsilon_0 is the . For (\epsilon \approx 78.5), this yields large negative values for small, highly charged ions, reflecting strong stabilization from the high screening. Despite its simplicity, the model has notable limitations: it assumes a uniform , neglecting molecular-scale details like solvent structure and solute size effects beyond the radius r, and it overestimates solvation energies for hard ions by ignoring short-range repulsions and specific ion-solvent orientations. Extensions such as the -Kirkwood model incorporate solute , accounting for induced dipoles that enhance electrostatic interactions in polar solvents. Representative examples illustrate these trends: the absolute solvation free energy of in is approximately -529 /, significantly more negative than that of at -306 /, due to 's smaller amplifying the $1/r dependence in the Born expression. Solvation energies are conventionally reported in /, with absolute values challenging to measure directly and often derived from thermodynamic cycles, while relative values between ions are more readily obtained experimentally.

Enthalpy, Entropy, and Free Energy

The of solvation, \Delta G_\text{solv}, quantifies the net stabilization of a solute upon transfer from the gas phase to and is given by the relation \Delta G_\text{solv} = \Delta H_\text{solv} - T \Delta S_\text{solv}, where \Delta H_\text{solv} is the solvation , \Delta S_\text{solv} is the solvation , and T is the . This thermodynamic framework, rooted in the transfer of a solute to a fixed position in the solvent, determines the spontaneity and extent of solvation, with negative \Delta G_\text{solv} indicating favorable . The solvation enthalpy \Delta H_\text{solv} arises from competing contributions: an endothermic term associated with cavity formation, which disrupts solvent-solvent interactions to create space for the solute, and an exothermic term from attractive solute-solvent interactions, such as van der Waals forces or hydrogen bonding. In polar solvents like water, the exothermic interactions often dominate for charged or polar solutes, leading to overall negative \Delta H_\text{solv}, while for nonpolar solutes, the cavity cost can make \Delta H_\text{solv} less favorable. Solvation \Delta S_\text{solv} typically decreases (negative \Delta S_\text{solv}) due to the ordering of molecules in the around the solute, restricting their translational and rotational freedom. Conversely, for hydrophobic solutes in , \Delta S_\text{solv} can become positive as the process releases structured from clathrate-like cages, increasing ; this gain drives the in amphiphilic solvation, where nonpolar groups aggregate to minimize surface exposure. dependence varies by : in aqueous media, the intensifies with rising T due to the amplified -T \Delta S_\text{solv} term, whereas solvents often exhibit more enthalpy-dominated solvation with weaker sensitivity. Enthalpy-entropy compensation is a pervasive feature in solvation, where favorable enthalpic changes (more negative \Delta H_\text{solv}) are offset by unfavorable entropic changes (less positive or more negative \Delta S_\text{solv}), resulting in relatively invariant \Delta G_\text{solv}. This correlation stems from solvent-mediated effects, such as adjustments in structure that couple energetic and configurational penalties. In ion-specific contexts, the ranks ions by their influence on solvation and protein stability: ions (e.g., SO_4^{2-}) promote enthalpic stabilization via enhanced water structuring, while chaotropic ions (e.g., SCN^-) boost by loosening shells, thereby modulating protein unfolding free energies.

Applications in Complex Systems

Macromolecules and Biomolecules

In proteins, the hydration shell plays a critical role in stabilizing secondary structures such as alpha-helices and beta-sheets by forming hydrogen bonds with polar backbone and side-chain groups, thereby modulating local conformational dynamics and preventing aggregation. This shell consists of bound water molecules that are tightly associated with the protein surface through specific interactions, contrasting with free bulk water that exhibits faster rotational and translational motion. Vicinal water within this shell displays altered physicochemical properties compared to bulk solvent, including higher viscosity due to restricted hydrogen bonding networks influenced by hydrophilic protein surfaces. Solvation extends by emphasizing that while the sequence dictates the native fold, the solvent environment drives the folding pathway through water expulsion from the hydrophobic core, achieving most structural formation prior to complete desolvation. For instance, burial of the hydrophobic core during folding minimizes unfavorable water interactions, enhancing thermodynamic stability via the . In enzyme active sites, solvation is finely tuned by partial desolvation and coordination with protein residues, enabling selective binding and , as seen in ion-dipole interactions that replace bulk water solvation shells. Nucleic acids, particularly DNA, rely on groove hydration to influence flexibility and molecular recognition. The minor groove, rich in electronegative atoms, accommodates ordered water networks that spine the helix and restrict bending, thereby modulating DNA curvature and protein-binding affinity. Sequence-specific hydration patterns in the major and minor grooves affect deformability; for example, A-tract regions exhibit enhanced minor groove solvation that promotes straight conformations essential for regulatory protein recognition. These hydration layers also facilitate indirect readout mechanisms, where water-mediated hydrogen bonds transmit sequence-dependent signals to binding partners. Key solvation concepts in biomolecular interactions include water bridges that link and residues, stabilizing transition states through dynamic hydrogen-bonding networks that enhance specificity and rate. Dehydration penalties arise during binding when removing ordered from interfaces incurs an energetic cost, often offset by favorable direct interactions but critical for predictions in protein-ligand complexes.

Supramolecular Assemblies

Supramolecular represent organized structures formed through non-covalent interactions where plays a pivotal role in stabilizing the architecture and driving . In these systems, molecules interact differentially with molecular components, influencing dynamics and functionality. For instance, amphiphilic molecules aggregate to minimize unfavorable contacts, creating distinct solvation environments that dictate the overall . In micelles and vesicles, solvation drives the of amphiphiles by promoting the inward orientation of hydrophobic tails to evade , while polar heads remain solvated at the . This reduces the penalty associated with exposing nonpolar regions to the aqueous , leading to spherical micelles at low concentrations and bilayer vesicles at higher ones. Electrostatic and hydrogen-bonding interactions between polar headgroups and molecules further stabilize these structures, preventing collapse and enabling dynamic transitions, such as from micelles to vesicles upon changes in or concentration. The (CMC), the threshold level for formation, is strongly influenced by properties, with polar solvents like lowering CMC through enhanced hydrophobic interactions compared to mixed aqueous-organic systems. In water-ethylene glycol mixtures, for example, increasing the organic fraction raises the CMC of by disrupting 's structured layer, thereby weakening the driving force for assembly. This solvent-dependent CMC highlights how solvation modulates the balance between gain from water release and enthalpic costs in supramolecular organization. Host-guest chemistry exemplifies solvation's role in selective binding, as seen in cyclodextrins where hydrophobic cavities encapsulate guests, driven by differential solvation energies that favor desolvation of the guest upon complexation. Binding affinities in β-cyclodextrin systems, calculated via and end-point free energy methods like MM/PBSA, reveal that solvation contributions account for significant portions of the change, with root-mean-square errors around 2.2 kcal/mol when corrections are included. This desolvation process enhances stability in aqueous media, enabling applications in molecular recognition. Reverse micelles illustrate solvation in nonpolar solvents, where form inverted structures with polar heads inward, solubilizing water or polar molecules in the core while hydrophobic tails interact with the apolar medium. Polar solvents trigger stable reverse micelle formation by swelling the core, leading to spherical aggregates with sizes tunable by and tail length; for instance, simulations show average diameters remaining constant despite thermal contraction due to fixed polar core volumes. Similarly, solvation shells around nanoparticles consist of layered solvent molecules that enhance colloidal , with porous exhibiting extended shells that penetrate internal cavities, influencing assembly and preventing aggregation through solvophobic interactions. Entropy-driven assembly is a key concept in these systems, where the release of solvating molecules from hydrophilic segments increases overall , compensating for unfavorable enthalpic changes. In amphiphilic bisimides with oligoethylene glycol dendrons, yields negative free energies (e.g., -21.9 kJ/mol) primarily from entropic gains of +33.1 J/mol·K, as shells around glycol units are disrupted during aggregation. This mechanism parallels biomolecular processes but is distinct in its emphasis on aggregate-level solvation dynamics. In , liposome hydration exemplifies solvation's practical impact, with thin hydration shells of 6-7 water molecules around phosphocholine headgroups controlling mobility and . These shells act as lubricants, enabling diffusion coefficients up to sixfold higher in hydrated states, which is crucial for encapsulating and releasing therapeutics; increases activation energies by twofold, allowing tunable for targeted without compromising upon rehydration. Solvation in the bilayer and headgroup regions further dictates partitioning, with hydrophobic drugs favoring the solvophobic interior.

Methods of Study

Experimental Techniques

Experimental techniques play a crucial role in elucidating the structural and dynamic aspects of solvation, providing direct insights into solvent-solute interactions without relying on theoretical models. These methods span a range of timescales and length scales, from ultrafast processes to properties, enabling researchers to quantify solvation shells, hydrogen bonding, and changes in various systems. Key approaches include spectroscopic, calorimetric, and techniques, which reveal how solvents reorganize around solutes like ions, small molecules, and biomolecules. Nuclear magnetic resonance (NMR) spectroscopy is widely used to investigate solvation dynamics, particularly water exchange rates in the hydration shells of proteins and ions. By measuring relaxation dispersion and chemical exchange, NMR can detect slowing of water exchange from the immediate vicinity of a protein surface to larger distances, with rates typically on the order of microseconds to milliseconds. For instance, in protein hydration studies, water molecules in the first solvation layer exhibit residence times of 10-100 ns, as determined by proton transverse relaxation rates. This technique has been instrumental in mapping dynamical hydration layers around macromolecules, highlighting how surface topology influences water mobility. Infrared (IR) and probe shifts associated with solvation by detecting vibrational frequency changes in solvent-solute bonds. These methods reveal redshifts in O-H stretching modes of upon hydrogen bonding to solutes, with shifts up to 200 cm⁻¹ observed in aqueous solutions of polar molecules. In hydrogen-bonded systems, IR and Raman spectra show characteristic intensity and frequency alterations that reflect solvation-induced perturbations in the solvent network, allowing quantification of bond strengths and orientations. For example, in alcohol-water mixtures, Raman shifts in the 3200-3600 cm⁻¹ region indicate progressive disruption of water structure by hydrophobic groups. Isothermal titration calorimetry (ITC) measures the enthalpy change (ΔH) of solvation by quantifying heat released or absorbed during solute-solvent interactions in solution. This technique titrates a solute into a solvent reservoir, yielding enthalpies for or processes that include solvation contributions, typically ranging from -10 to -50 kcal/mol for non-covalent molecular interactions such as protein-ligand . ITC provides thermodynamic parameters like directly, distinguishing solvation effects from other contributions in processes such as protein-ligand , where desolvation penalties can dominate . Dielectric relaxation spectroscopy assesses solvent reorientation times around solutes by monitoring the frequency-dependent dielectric response in the GHz to THz . This method detects slowed reorientation of dipoles in solvation shells, with relaxation times increasing from ~8 ps in to 20-50 ps near ions due to restricted motion. In solutions, it reveals ion-specific effects, such as enhanced structuring around chaotropes versus kosmotropes, through shifts in the Debye relaxation peak. Solvatochromic dyes serve as polarity probes for solvation environments by exhibiting shifts in or spectra proportional to . These dyes, such as Reichardt's betaine, quantify local via the solvatochromic parameter E_T, which correlates with dielectric constant and hydrogen-bonding ability, enabling assessment of microheterogeneity in mixed solvents or near biomolecules. For instance, in aqueous s, dyes reveal preferential solvation by over the component, with shifts up to 100 nm. Extended X-ray absorption fine structure (EXAFS) spectroscopy determines the structure of solvation shells by analyzing oscillations in absorption spectra beyond the edge. This resolves first-shell coordination numbers and lengths, such as approximately 6 oxygen atoms at 2.4 for aqueous Na⁺, providing atomic-level details on geometry. In halide solutions, EXAFS confirms asymmetric solvation with varying shell thicknesses, distinguishing contact ion pairs from fully solvated s. Fluorescence correlation spectroscopy (FCS) examines solvation in biomolecules by tracking fluorescence fluctuations to infer local diffusion and hydration dynamics. In protein systems, FCS measures anomalous diffusion due to hydration layer viscosity, with diffusion coefficients reduced by 20-50% compared to bulk solvent, reflecting coupled water-biomolecule motions on nanosecond scales. This approach highlights how hydration shells modulate conformational dynamics in enzymes and DNA. Recent advances in terahertz (THz) spectroscopy have enabled probing of ultrafast solvation dynamics, capturing collective solvent motions on picosecond timescales post-2020. THz time-domain spectroscopy reveals bimodal relaxation in ion solutions, with fast (~0.2 ps) intramolecular modes and slower (~1-10 ps) intermolecular reorientations influenced by solvation shells. For example, in alkali halide solutions, THz-Raman studies show cations weakening hydrogen bonds between first and second water layers, altering dielectric responses by up to 30%. These developments provide unprecedented resolution of initial solvation stages, linking to thermodynamic parameters like free energy changes in complex environments.

Computational Simulations

Computational simulations play a crucial role in modeling solvation by predicting solvent-solute interactions at the level, offering insights into dynamic processes that complement experimental . Explicit solvent models treat the as molecules, explicitly including all atoms in the simulation to capture detailed interactions such as hydrogen bonding and hydrophobic effects. In (MD) simulations using all-atom explicit , the system evolves according to classical , allowing for the study of time-dependent solvation phenomena like and conformational changes influenced by the surrounding medium. (MC) simulations, on the other hand, focus on properties by randomly sampling configurations according to the , providing accurate solvation free energies through without time evolution. These explicit approaches are computationally intensive but essential for systems where structure is critical, such as in biomolecular environments. Force fields are paramount in explicit simulations, parameterizing to represent solvent behavior accurately. The TIP3P model, a three-site rigid , is widely used in MD simulations for its balance of computational efficiency and fidelity in reproducing water's dielectric properties and hydrogen bonding in solvation contexts. However, simulations conducted in vacuo—without explicit —severely limit accuracy by neglecting solvent-mediated forces, leading to unrealistic solute conformations and overestimated intramolecular interactions that fail to capture solvation's stabilizing or destabilizing effects. Implicit solvent models approximate the solvent as a continuous medium, reducing computational cost while accounting for average solvation effects, making them suitable for large-scale or long-time simulations. The Polarizable Continuum Model (PCM) embeds the solute in a cavity within a polarizable , solving to compute reaction field energies that polarize the solute and yield solvation free energies. Similarly, the Generalized Born Surface Area (GBSA) model approximates electrostatic solvation via a generalized for interactions, augmented by a surface area term for nonpolar contributions, enabling efficient binding affinity predictions in protein-ligand complexes. Recent advances have integrated (ML) to enhance simulation accuracy and speed, particularly for solvation free energies. Neural network-based ML potentials, trained on quantum mechanical data, approximate surfaces for explicit solvent systems, achieving near-ab initio accuracy for solvation at a fraction of the computational cost, as demonstrated in predictions for molecules in aqueous environments post-2020. As of 2025, further developments include the Solvation Free Energy Path Reweighting (ReSolv) framework and hybrid /molecular mechanics (ML/MM) interfaces, which improve accuracy and stability for free energy calculations in diverse solvent systems. Enhanced sampling techniques like address ergodic sampling limitations by adding history-dependent bias potentials to collective variables, accelerating exploration of solvation landscapes in MD simulations of protein adsorption and solvent reorganization. These methods find application in complex systems, such as simulating protein shells, where reveals slowed water and structured layers extending up to 10 from the surface, influencing protein stability and function. In ionic liquids, explicit and MC simulations elucidate solvation structures around biomolecules, showing how ions form competitive layers that modulate and compared to traditional solvents.

References

  1. [1]
    12.1 The Dissolution Process – Chemistry Fundamentals
    A solution forms when two or more substances combine physically to yield a mixture that is homogeneous at the molecular level. The solvent is the most ...
  2. [2]
    13.1 Factors Affecting Solution Formation
    Solvation is the process in which solute particles are surrounded by solvent molecules. When the solvent is water, the process is called hydration. The ...Missing: definition | Show results with:definition
  3. [3]
    13.3: Solvation Processes
    ### Summary of Solvation Processes
  4. [4]
    Solvation Thermodynamics and Its Applications - PMC - NIH
    “When a solvent and a solute molecule link together with weak bonds, the process is called solvation”. In Gurney's classical book (1953) [2], we find the ...
  5. [5]
    CH150: Chapter 7 - Solutions - Chemistry - Western Oregon University
    This process is called solvation and is illustrated in Figure 7.2. When the solvent is water, the word hydration, rather than solvation, is used. In general ...
  6. [6]
    7.1: Solutions – Homogeneous Mixtures - Maricopa Open Digital Press
    The interactions between the solute particles and the solvent molecules is called solvation. A solvated ion or molecule is surrounded by solvent molecules ( ...<|control11|><|separator|>
  7. [7]
    Chemical Reactivity - MSU chemistry
    ... primary solvation shell about Na(+) and Cl(–) are shown here. The water dipoles are drawn as red arrows, and partial charges are noted. Additional water ...
  8. [8]
    [PDF] Solvation of Magnesium Dication: Molecular Dynamics Simulation ...
    Additionally, we will define a solvent-shared ion pair as a cation and anion separated by a water molecule that is in the first solvation shell of both ions, ...
  9. [9]
    [PDF] INTERACTION OF WATER WITH THE - OhioLINK ETD Center
    Primary solvation shell: both cations and anions have primarily solvation shells consisting of solvent molecules interacting directly with the ions b ...
  10. [10]
    [PDF] Development of the theory of electrolytic dissociation - Nobel Prize
    The question now arose in what respect the active state of the electrolytes differs from the inactive state. I gave an answer to this question in 1887. At.Missing: solvation | Show results with:solvation
  11. [11]
    Biochemistry, Dissolution and Solubility - StatPearls - NCBI Bookshelf
    Dissolution is the process where a solute in a gaseous, liquid, or solid phase dissolves in a solvent to form a solution.[1][2][3] Solubility is the maximum ...
  12. [12]
    solvation (S05747) - IUPAC Gold Book
    Any stabilizing interaction of a solute (or solute moiety ) and the solvent or a similar interaction of solvent with groups of an insoluble material.
  13. [13]
    solubility (S05740) - IUPAC Gold Book
    The analytical composition of a saturated solution, expressed in terms of the proportion of a designated solute in a designated solvent, is the solubility ...
  14. [14]
    [PDF] glossary of terms related to solubility | iupac
    Dec 28, 2016 · This glossary defines 166 terms used to describe solubility and related phenomena. The definitions are consistent with one another and with ...
  15. [15]
    Revisiting the hydration structure of aqueous Na+ - AIP Publishing
    Feb 27, 2017 · The measured Na–O coordination number from XRD is 5.5 ± 0.3. The measured structure is compared with both classical and first-principles density ...
  16. [16]
    6.6: Lattice Energy and Solubility - Chemistry LibreTexts
    Aug 4, 2025 · For a solid to be soluble, the energy required to break the lattice must be offset by the solvation of the ions or molecules in solvent.
  17. [17]
    [PDF] solute-solvent-interaction.pdf - Walsh Medical Media
    Jul 29, 2021 · Intermolecular interactions such as hydrogen bonding, ion-dipole interactions, and van der Waals forces all play a role in solvation.
  18. [18]
    Solvation-Driven Charge Transfer and Localization in Metal ...
    Apr 22, 2015 · In charge transfer (CT) processes in polar solvents, the response of the solvent always assists the formation of charge separation states by stabilizing the ...Introduction · Constrained Solvation-Driven... · Solvent-Driven Charge...
  19. [19]
    Water Dynamics in the Hydration Shells of Biomolecules
    in the collapse of the protein chain during folding through hydrophobic collapse and mediates binding through the hydrogen bond in complex formation. Water ...
  20. [20]
    The effects of charge transfer on the aqueous solvation of ions
    Jul 31, 2012 · Charges of solvated ions are not required to be integers and water molecules can become charged, either through charge transfer with ions or ...
  21. [21]
    Origin of hydrophobicity and enhanced water hydrogen bond ...
    Dec 27, 2016 · We present unequivocal experimental proof for strengthened water hydrogen bonds near purely hydrophobic solutes, matching those in ice and clathrates.
  22. [22]
    Ion Pairing | Chemical Reviews - ACS Publications
    The formation of ion pairs is strongly influenced by the solvation of the ions; hence, the transfer of ion pairs between solvents of different solvation ...Theoretical Treatments of Ion... · Thermodynamic... · Solvation and Ion Pairing
  23. [23]
    Polar Protic Solvents - LON-CAPA OCHem
    Polar protic solvents have a hydrogen atom attached to an electronegative atom, usually oxygen, and are represented by the formula ROH. Examples include water, ...
  24. [24]
    Chapter 8 Notes
    Protic solvent: a solvent that is a hydrogen bond donor · Aprotic solvent: a solvent that cannot serve as a hydrogen bond donor · Solvents are classified as polar ...
  25. [25]
    [PDF] Marine Organic Geochemistry
    Chemists classify solvents according to their polarity: • polar protic • dipolar aprotic • non-polar. Polar Protic Solvents • Protic refers to a hydrogen atom ...
  26. [26]
    Static dielectric constant of pure water at 25 °C - BioNumbers
    The static dielectric constant of pure water is 78.4 at 25 °C (primary source), but dissolved ions can decrease this value substantially due to the tendency of ...
  27. [27]
    Empirical parameters for donor and acceptor properties of solvents
    Two empirical parameters are used namely the donicity or donor number (DN) to describe the nucleophilic behaviour and the acceptor number (AN) to describe ...
  28. [28]
    The acceptor number — A quantitative empirical parameter for the ...
    The Acceptor Number (AN) characterizes the electrophilic properties of solvents and is derived from 31P-n.m.r. measurements. It helps choose appropriate  ...
  29. [29]
    Quantitative Measures of Solvent Polarity | Chemical Reviews
    Factors i−iv are collectively well-known as “solvent effects” and are conventionally ascribed to a solvent's “polarity”. The term “polarity” is usually related ...<|control11|><|separator|>
  30. [30]
    Solvation properties of protic ionic liquids and molecular solvents
    Dec 2, 2019 · In this study, we have investigated the solvation properties of 12 protic ionic liquids (PILs) and 9 molecular solvents based on the Kamlet–Abboud–Taft′ (KAT) ...
  31. [31]
    Polar Attributes of Supercritical Carbon Dioxide - ACS Publications
    Historically CO2 was treated as a nonpolar solvent, primarily because of its low dielectric constant and zero molecular dipole moment. CO2 has also been ...
  32. [32]
    A salient effect of density on the dynamics of nonaqueous electrolytes
    Apr 24, 2017 · Our study finds that a small variation in density can induce a significant effect on the mobility of electrolytes but does not influence the solvation ...
  33. [33]
    How does the solvent composition influence the transport properties ...
    Apr 8, 2021 · As a function of the solvent composition, the following properties were examined: the density, viscosity, ionic conductivity, solvation states ...
  34. [34]
    Quantum Mechanical Continuum Solvation Models - ACS Publications
    The systematic calculations of the solvation free energy of ions date back to Born's seminal paper. ... Theory and applications of the generalized Born solvation ...
  35. [35]
    A molecular Debye-Hückel theory of solvation in polar fluids
    However, the Born model has several limitations. First, the Born model overestimates the solvation energy as long as a is interpreted as the cavity size of the ...INTRODUCTION · The electric potential of a... · The dielectric response...
  36. [36]
    Absolute ion hydration enthalpies and the role of volume within ...
    May 25, 2017 · In 2006 Kelly, Cramer and Truhlar were able to confirm the absolute aqueous solvation free energy of the proton, ΔhydGo(H+,g) as −1112.5 kJ mol− ...
  37. [37]
    Solvation Thermodynamics and Its Applications - MDPI
    In this article, we start by describing a few “definitions” of the solvation processes, which were used in the literature until about 1980.
  38. [38]
    Practical Aspects of Free-Energy Calculations: A Review
    By inverting the charge distribution in the artificial solutes, large differences in the hydration free energy of up to 40 kJ mol–1 were found, largely driven ...
  39. [39]
    A View of the Hydrophobic Effect | The Journal of Physical Chemistry B
    At low temperatures, thermodynamic processes are driven to lower their enthalpies, and at high temperatures, they are driven to states of high entropy.
  40. [40]
    Enthalpy–entropy compensation: the role of solvation
    Oct 28, 2016 · We review long-established and recent cases of EEC and conclude that the large fluctuations in enthalpy and entropy observed are too great to be ...
  41. [41]
    Entropy−Enthalpy Compensation in Solvation and Ligand Binding ...
    These metrics are regularly updated to reflect usage leading up to the last few days. Citations are the number of other articles citing this article, calculated ...
  42. [42]
    Protein Stabilization and the Hofmeister Effect: The Role of ... - NIH
    Our results are in reasonable agreement with the idea that Hofmeister anions increase ProtL stability by changing the surface tension of the solution.
  43. [43]
    Water Dynamics in Protein Hydration Shells: The Molecular Origins ...
    We show that all four proteins have very similar hydration shell dynamics, despite their wide range of sizes and functions, and differing secondary structures.
  44. [44]
    Hydrophilicity and the Viscosity of Interfacial Water - ResearchGate
    Aug 9, 2025 · Viscosity in interfacial water is increased by hydrophilicity and reduced by hydrophobicity, implying the strength of the hydrogen bond is ...
  45. [45]
    Protein folding mediated by solvation: Water expulsion and ... - PNAS
    These results support a folding mechanism where most of the structural formation of the protein is achieved before water is expelled from the hydrophobic core.
  46. [46]
    Quantitative theory of hydrophobic effect as a driving force of protein ...
    Various studies suggest that the hydrophobic effect plays a major role in driving the folding of proteins. In the past, however, it has been challenging to ...
  47. [47]
    Ion–dipole interactions and their functions in proteins - Sippel - 2015
    Apr 11, 2015 · 2-4 These solvation energies arise from the classic ion–dipole interaction between ions and water dipoles. In some cases the unbound protein ...Ion--Dipole Definition · Ion--Dipoles In Nature · Ion Channel Selectivity...
  48. [48]
    Specific minor groove solvation is a crucial determinant of DNA ...
    Nov 27, 2014 · Here we show that the differential abilities of various DNA sequences to support formation of a highly ordered and stable minor groove solvation ...
  49. [49]
    Sequence Dependencies of DNA Deformability and Hydration in the ...
    DNA deformability and hydration are both sequence-dependent and are essential in specific DNA sequence recognition by proteins.
  50. [50]
    Water at DNA surfaces: Ultrafast dynamics in minor groove recognition
    Water molecules at the surface of DNA are critical to its equilibrium structure, DNA–protein function, and DNA–ligand recognition.
  51. [51]
    Water Determines the Structure and Dynamics of Proteins - PMC
    The two interactions are coupled dynamically, leading to a water-mediated induced-fit between the substrate and the enzyme in the enzyme catalysis, for instance ...
  52. [52]
    Solvation Contribution to the Free Energy of Ligand Binding is ...
    Feb 12, 2021 · We introduce a computational method to calculate the desolvation penalty upon ligand binding based on the analysis of the surface hydration structure.
  53. [53]
    Self-Assembly of Amphiphiles into Vesicles and Fibrils
    Therefore, they have a tendency to desorb upon heating, and this induces the change in the molecular geometry from vesicles to wormlike micelles. Figure 4.
  54. [54]
    Molecular forces in the self-organization of amphiphiles
    ... micelles to vesicles. Colloid and Polymer Science 2015, 293 (12) , 3545-3554 ... Micelles upon Changing the External Solvent Composition and Their Impact on Gold ...
  55. [55]
    Surfactant Self-Assembling and Critical Micelle Concentration
    May 6, 2020 · Critical micelle concentration (CMC) is the main chemical–physical parameter to be determined for pure surfactants for their ...
  56. [56]
  57. [57]
    Polar Solvents Trigger Formation of Reverse Micelles
    ### Summary of Reverse Micelles Formation and Solvation Properties
  58. [58]
    Solvation of Nanoscale Materials - ACS Publications
    Aug 14, 2024 · Here, we discuss classical models of solvation and colloidal stability for nonporous and pseudoporous (proteins and polymers) materials as a basis.3. Solvation At... · Figure 3 · 4.2. Porous Frameworks
  59. [59]
    Thermodynamic insights into the entropically driven self-assembly of ...
    Aug 20, 2019 · The entropically favoured nature of the self-assembly is attributed to the release of water molecules from the glycol units which is ...
  60. [60]
    Hydration Layer of Only a Few Molecules Controls Lipid Mobility in ...
    Aug 3, 2021 · We introduce a new method of obtaining dehydrated, phase-separated, supported lipid bilayers (SLBs) solely by controlling the decrease of their environment's ...
  61. [61]
    Structure-Based Prediction of Drug Distribution Across the ...
    Sep 1, 2014 · Solvation of drugs in the core (C) and headgroup (H) strata of phospholipid bilayers affects their physiological transport rates and ...
  62. [62]
    Comparison of Implicit and Explicit Solvent Models for the ...
    Feb 28, 2017 · Here, we evaluate the predictive power of implicit solvent models for solvation free energy of organic molecules in organic solvents.Introduction · Methods · Results · References
  63. [63]
    Absolute free energy of solvation from Monte Carlo simulations ...
    Absolute free energy of solvation from Monte Carlo simulations using combined quantum and molecular mechanical potentials | The Journal of Physical Chemistry.
  64. [64]
    Force fields for monovalent and divalent metal cations in TIP3P ...
    Feb 21, 2018 · Here, we develop force field parameters for the metal cations Li + , Na + , K + , Cs + , Mg 2+ , Ca 2+ , Sr 2+ , and Ba 2+ in combination with the TIP3P water
  65. [65]
  66. [66]
    Polarizable continuum models for quantum-mechanical descriptions
    Apr 22, 2016 · Polarizable continuum solvation models are nowadays the most popular approach to describe solvent effects in the context of quantum mechanical calculations.INTRODUCTION · II. TOWARDS VERY LARGE... · IV. BEYOND THE STANDARD...
  67. [67]
    Generalized Born Implicit Solvent Models for Biomolecules - PMC
    Jul 22, 2019 · The basic idea of implicit solvation is to integrate out explicit solvent degrees of freedom, incorporating their thermodynamic effects into a ...
  68. [68]
    Improved prediction of solvation free energies by machine-learning ...
    Jun 18, 2021 · We develop and introduce the Machine-Learning Polarizable Continuum solvation Model (ML-PCM) for a substantial improvement of the predictability of solvation ...
  69. [69]
    Accelerating Solvent Dynamics with Replica Exchange for Improved ...
    Oct 21, 2023 · Coupling MD simulations with metadynamics (enhanced sampling) has fast become a common method for sampling the adsorption of such proteins.
  70. [70]
    An extended dynamical hydration shell around proteins - PNAS
    Molecular dynamics simulations indicate water dynamics in the solvation layer around one protein to be distinct from bulk water out to ≈10 Å. At higher protein ...
  71. [71]
    Computational solvation analysis of biomolecules in aqueous ionic ...
    Here, we report on current computational approaches to characterize the impact of the ionic liquid ions on the structure and dynamics of the biomolecule and ...