Fact-checked by Grok 2 weeks ago

Solvation shell

A solvation shell is the layer of molecules that closely surround and interact with a solute particle in a , forming a structured region that stabilizes the solute through electrostatic, bonding, and other intermolecular forces. This shell, often the first of multiple concentric layers, consists of molecules acting as nearest neighbors to the solute, with the number of such molecules defined by the solute's . In aqueous environments, it is specifically termed the hydration shell, where molecules orient their dipoles toward charged or polar solutes, such as ions, to minimize the system's . The structure of the solvation shell is typically characterized using the (RDF), g(r), which describes the probability density of finding a at a distance r from the solute; the first solvation shell is delimited by the position of the first minimum in this function, often around 3–4 for small ions in . For example, monovalent ions like Na^+ exhibit a of approximately 5–6 molecules in their primary shell, while divalent ions like Mg^{2+} coordinate 6, forming octahedral arrangements. Beyond the first shell lies the second solvation shell, where organization is less pronounced but still influenced by the solute, transitioning gradually to bulk behavior. These structural features vary with solute size, charge, and type, impacting properties like and reactivity. Dynamically, the solvation shell is not static but undergoes continuous fluctuations and exchange of molecules, with residence times ranging from picoseconds for non-hydrogen-bonded interactions to nanoseconds in strongly coordinating cases. exchange rates, quantified by the number of continuous neighbors (CNS), reveal slower for multiply charged species due to stronger binding energies. This motion governs key processes such as mobility, proton transfer in (e.g., via involving shell reorganization), and the activation barriers for chemical reactions in solution. In non-aqueous , similar principles apply, but shell composition can include mixed , leading to preferential solvation where one component dominates the shell. The plays a central role in thermodynamic properties, contributing significantly to the of , which determines and phase behavior. For instance, the enthalpic stabilization from shell formation often competes with entropic costs from solvent ordering, explaining phenomena like the where nonpolar solutes induce cage-like ordering in the surrounding molecules. Experimental techniques such as , neutron scattering, and vibrational spectroscopy, complemented by simulations, have elucidated these shells in diverse systems, from simple to biomolecules. Understanding solvation shells is essential for fields like , where they influence performance, and biochemistry, where they modulate and .

Fundamentals

Definition and Formation

A solvation shell refers to the oriented layer of molecules that surrounds a solute , , or particle in a , arising primarily from electrostatic interactions such as ion-dipole forces, as well as hydrogen bonding and van der Waals attractions. This structured arrangement stabilizes the solute by compensating for its exposure to the bulk environment. The formation of the solvation begins with the primary , which constitutes the first coordination layer of solvent molecules in direct, strong contact with the solute through specific or electrostatic interactions. Beyond this, the secondary emerges via weaker, indirect interactions between the primary and adjacent solvent molecules, creating a of influence that diminishes with distance from the solute. The overall is driven by the minimization of , where solvent molecules reorient to accommodate the solute's of the structure. The concept of solvation shells was first systematically described in the early , notably in the context of ion , by Bernal and Fowler in their 1933 theory of and ionic solutions. Key factors influencing shell formation include the solute's size, , and the solvent's polarity. Smaller, highly charged solutes, such as Li⁺, induce tighter primary shells due to stronger electrostatic fields, while larger or less charged species like K⁺ result in more diffuse arrangements. In polar solvents like , dipole moments enhance ion-dipole coupling, promoting stable shells; nonpolar solvents, by contrast, yield weaker or absent structuring around charged solutes. For instance, in , the Na⁺ ion typically coordinates 5–6 molecules in its primary hydration shell, forming a roughly octahedral that reflects the balance between and solvent accessibility.

Structure and Dynamics

The structure of the solvation shell is characterized by distinct layers of molecules organized around the solute, as revealed by radial distribution functions (RDFs). The RDF, which describes the probability of finding a atom at a given distance from the solute, typically exhibits a sharp first peak corresponding to the primary solvation shell at distances of approximately 2.0–3.5 for oxygen atoms around ions, indicating direct coordination. A second peak, often broader and less intense, appears at 4.0–5.5 , delineating the secondary shell where interactions are weaker and more influenced by the primary layer. The , defined as the average number of molecules in the primary shell obtained by integrating the RDF up to its first minimum, varies with solute type; for example, Li⁺ in has a of about 4.1, Na⁺ around 5.5, and K⁺ near 7.0. These structural features highlight a dense, ordered in the inner shell that transitions to bulk-like behavior beyond the second shell. Within the solvation shell, particularly in polar solvents like , solvent molecules exhibit pronounced orientational order due to interactions with the solute's . dipoles align radially around charged solutes, with the degree of alignment increasing with ionic , often quantified by the cosine of the dipole angle relative to the solute-solvent (e.g., crossing from random to ordered at cos θ ≈ 0.3). This alignment disrupts the isotropic tetrahedral structure of bulk , leading to reduced hydrogen bonding in the primary shell—where the average number of H-bonds per drops below the bulk value of ~3.5—and promotes directional networks that stabilize the shell. In shells around biomolecules or ions, this results in structured motifs like tetrahedral or octahedral arrangements, enhancing local order compared to the bulk solvent's dynamic . The dynamics of solvent molecules in the solvation shell involve characteristic timescales for residence and exchange that differ markedly from bulk solvent. In bulk water, the residence time—related to hydrogen bond lifetimes and reorientational motion—is on the order of 1–2 picoseconds, reflecting rapid molecular tumbling. In contrast, water in the primary solvation shell has longer residence times, typically 10–100 picoseconds or more, due to stronger solute-solvent interactions; for instance, around small cations like Na⁺, times are ~20–50 ps, while in confined sites near proteins or DNA, they extend to hundreds of picoseconds. Exchange rates between the shell and bulk, governed by diffusive jumps, are correspondingly slowed by factors of 2–6 relative to bulk, with some sites showing up to 20-fold retardation, as solvent molecules must overcome energetic barriers to escape the ordered shell. Several factors influence the and of the solvation shell, notably , which modulates shell thickness and properties. As increases, RDF peaks broaden slightly, indicating a minor expansion of shell thickness (e.g., first minimum shifting by ~0.1–0.2 from 298 K to 318 K), and coordination numbers decrease marginally due to enhanced —such as from ~6.2 to 5.9 for Na⁺ in DMF. In shell regions, is elevated compared to bulk solvent, impeding motion, while coefficients of solvent molecules are reduced (e.g., ~50–80% of bulk values for around cations), though still higher than the solute's diffusion, reflecting a semi-confined . These effects underscore the shell's role as a dynamic sensitive to thermal perturbations.

Theoretical Models

Continuum Models

Continuum models of the solvation shell treat the solvent as a continuous medium surrounding the solute, neglecting explicit molecular details of the solvent structure. These approaches emerged in the early to describe electrostatic interactions in solutions, providing a macroscopic for estimating solvation energies without resolving individual solvent molecules. The foundational continuum model is the Born solvation model, introduced by in 1920, which calculates the electrostatic of transferring an from to a solvent. In this model, the is represented as a charged sphere of r embedded in a dielectric continuum with constant \varepsilon, and the solvation \Delta G is given by \Delta G = -\frac{N_A z^2 e^2}{8 \pi \varepsilon_0 r} \left(1 - \frac{1}{\varepsilon}\right), where N_A is Avogadro's number, z is the ion charge number, e is the , and \varepsilon_0 is the . This equation arises from the difference in the ion's in (\varepsilon = 1) and in the , assuming the charge is uniformly distributed on the sphere's surface and the solvation shell forms an abrupt boundary. The model laid the groundwork for modern theories by linking solvation energetics to the solvent's properties, influencing subsequent developments in . An extension to polar, non-ionic solutes came with the Onsager reaction field model in 1936, which accounts for the interaction between a solute and the polarized . Here, the solute is placed in a spherical within the , inducing a reaction field that opposes the and provides a correction to the solvation free energy. This model improves upon the Born approach for molecules with permanent by incorporating orientational polarization effects, making it suitable for estimating solvation in polar like . Modern continuum models, such as the Polarizable Continuum Model (PCM), extend these ideas by solving the Poisson equation for the solute's electrostatic potential in a polarizable , accounting for cavity formation and charge distribution on the solute-solvent boundary. PCM and its variants, like Integral Equation Formalism PCM (IEF-PCM), provide more accurate solvation free energies for larger molecules by incorporating nonequilibrium effects and are widely used in quantum mechanical calculations. Despite their simplicity and utility, continuum models have significant limitations. They ignore the molecular structure of the , leading to inaccuracies for small ions where the solvation shell's atomic-scale organization dominates electrostatic screening. For instance, the model often overestimates solvation energies for small ions like ^+ to its assumption of a sharp , which fails to capture enhanced near the solute. Similarly, the Onsager model performs poorly for non-spherical solutes, as its spherical approximation neglects shape-dependent cavity formation and anisotropic . Additionally, these models tend to overestimate solvation effects in low-dielectric s, where molecular discreteness and cavity fluctuations become prominent.

Discrete and Molecular Models

Discrete and molecular models of the shell treat molecules as distinct entities, enabling detailed descriptions of local interactions and structures that approaches often overlook. These models emphasize atomistic representations to capture the discrete nature of , particularly for ions and small solutes where the first shell dominates energetic contributions. By accounting for individual molecular orientations and packing, they provide insights into shell formation and stability beyond macroscopic responses. One adsorption-like framework for ion is the Langmuir-inspired model, which analogizes the first coordination shell to a where molecules adsorb onto the solute surface with saturation limits. In this approach, the local in the shell is modeled using Langmuir isotherm parameters, such as maximum occupancy correlated with the solute's surface field, allowing predictions of coordination numbers in supercritical or dense fluids. For instance, in supercritical , this model relates the of to local in the shell to temperature-independent adsorption strengths, fitting experimental energies with deviations under 5 kJ/mol. Complementary to this, the Blum model within the mean spherical approximation treats the shell as an effective charged shell around ions, incorporating non-additive core exclusions to compute pair correlations and free energies for halides in , achieving accuracies of 1.5 kJ/mol against data. Quasi-lattice theories further describe distribution around solutes through on a discretized , partitioning the into sites occupied by or voids. These models, such as the quasi-lattice quasi-chemical , evaluate preferential by balancing solute-solvent and solvent-solvent interactions, predicting compositions in binary mixtures via mixing free energies. A notable example is the significant structure , which views the liquid as a quasi-lattice blending gas-like and solid-like fractions, applied to solutions to model domain-structured around ions like NaCl, where the influences and coefficients. Such frameworks highlight how occupancy modulates , contrasting with continuum estimates for small-scale systems. Integral equation theories, such as the Reference Interaction Site Model (RISM), provide a statistical approach to compute structures by solving Ornstein-Zernike equations with approximations for site-site correlations. RISM captures radial distribution functions and coordination numbers in the solvation shell for molecular solutes in liquids, bridging explicit and implicit descriptions, and is often coupled with for hybrid models. Detailed insights into solvation shell properties are obtained through computational simulations, as discussed in the simulation approaches section. For example, studies reveal first hydration shell thicknesses around 3–5 for ions in , with coordination numbers of 5–6 for Na⁺ and slower exchange dynamics compared to bulk solvent. Advanced applications extend these models to complex interactions, such as water-mediated ion pairing and anisotropic solvation around proteins.

Applications in Solutions

Electrolyte Solutions

In electrolyte solutions, the solvation shell surrounding ionic solutes plays a crucial role in modulating electrostatic interactions and overall solution behavior. Unlike neutral solutes, experience strong orientational ordering of dipoles in their primary solvation shell, which screens the bare charge and influences thermodynamic properties such as activity coefficients. This shell formation is driven by ion-dipole attractions, leading to a structured layer of molecules that extends into secondary shells, affecting mobility and interionic distances. The solvation shell provides a short-range correction to the classical Debye-Hückel (DH) theory, which primarily accounts for long-range electrostatics in dilute solutions. In the extended DH framework, the mean ionic is approximated as \log_{10} \gamma \approx -\frac{0.51 z^2 \sqrt{I}}{1 + \sqrt{I}} (for at 25°C), where z is the charge and I is the ; however, deviations at higher concentrations arise from finite ion sizes and effects, incorporated via effective hydrated diameters or solvation terms that adjust for ion-solvent interactions. These corrections improve predictions of up to ~1 M by modeling the solvation shell's contribution to the excess , as seen in optimized models for various salts. The solvation number, defined as the average number of solvent ligands in the primary solvation shell of an ion, quantifies the extent of this immediate coordination; for example, the Li^+ ion in water typically coordinates six water molecules, forming an octahedral structure confirmed by neutron scattering. This number is determined experimentally from transport properties like ionic conductivity, where lower mobility of strongly solvated ions (e.g., Li^+ < Na^+) reflects the drag from the shell, or from solubility data, which correlate with hydration energies influencing salt dissolution equilibria. Solvation numbers generally range from 4 to 8 for monovalent ions in aqueous solutions, decreasing with ion size due to reduced charge density. Ion pairing in electrolyte solutions is profoundly influenced by solvation shell overlap, leading to distinct patterns. Inner-sphere complexes form when oppositely charged directly contact, with one replacing a in the other's primary shell, often in low-dielectric or for hard ; in contrast, outer-sphere complexes maintain separate shells separated by one or more molecules, favored in highly solvating media like where shell integrity prevents close approach. This distinction affects and reactivity, with shell overlap promoting pairing at higher concentrations and altering effective activities. The ranks ions by their specific effects on solution properties, attributable to variations in solvation shell strengths; chaotropic ions (e.g., I^-) form weaker, more polarizable shells that disrupt structure, while ions (e.g., SO_4^{2-}) create rigid, strongly bound shells that enhance water ordering. These differences manifest in ion-specific impacts on protein , where kosmotropes promote salting-out by stabilizing hydrophobic interfaces through indirect shell-mediated effects, without direct protein binding.

Non-Electrolyte Solutions

In non-electrolyte solutions, shells form around neutral solute molecules through short-range, non-electrostatic interactions, distinguishing them from the charge-dominated structures in solutions. These interactions include hydrogen bonding for polar nonelectrolytes, such as in , where the solute's functional groups directly coordinate with molecules, and dispersion forces for nonpolar solutes, leading to weaker and more diffuse shells. The structural characteristics, including numbers and the of in the shell, show linear dependencies on and concentration up to the complete solvation limit, reflecting the influence of the solute's on properties without significant differences from electrolytic behavior in basic volumetric terms. For hydrophobic nonelectrolytes in , the solvation shell often adopts a clathrate-like structure, as seen with , where molecules organize into a tetrahedral hydrogen-bonded network enclosing the solute, enhancing bond strengths comparable to those in with a of approximately 60 cm⁻¹ in the O-D stretching mode. This ordering, involving 10–15 ice-like hydrogen bonds per solute molecule, extends to nonpolar gases like O₂, whose low arises from hydrophobic effects that increase 's structural rigidity in the first shell without contracting O–O distances, supporting a moderate interpretation of the iceberg model over a rigid clathrate. In nonpolar organic solvents, such as solvating , the shell relies predominantly on dispersion forces due to comparable molecular polarizabilities, resulting in a loosely structured with minimal to solvent organization and near-ideal mixing behavior. In mixed solvent systems, preferential occurs when the composition deviates from the bulk, driven by differential affinities; for instance, in - mixtures, nonpolar nonelectrolytes like CO₂ preferentially accumulate in their at low concentrations, as evidenced by parameters that vary with temperature. This local enrichment is quantified via Kirkwood-Buff integrals (G_{ij}), which relate radial distribution functions to local mole fractions and composition fluctuations, providing insights into size mismatches and deviations from ideality in binary hydrocarbon or mixtures. Thermodynamic implications include enthalpy contributions from specific interactions like enhanced hydrogen bonding and penalties from ordering, which govern ; for hydrophobic cases, the negative from structured dominates, while dispersion-driven in nonpolar media favors enthalpic stabilization with smaller entropic costs. Kirkwood-Buff theory further connects these effects to fluctuations, enabling prediction of transfer free energies without assuming symmetrical ideality.

Biological and Macromolecular Contexts

Protein Hydration Shells

The hydration shell surrounding proteins consists of a vicinal water layer approximately 5-10 thick, where water molecules exhibit altered compared to . Within this layer, water can be classified as tightly bound, with residence times exceeding nanoseconds due to strong bonding with polar or charged protein residues, or diffusive, with shorter residence times on the order of picoseconds to tens of picoseconds, resembling behavior. This heterogeneity arises from the protein surface's chemical composition and , leading to a denser packing of molecules in the first , about 15% higher than in . The hydration shell plays crucial functional roles in maintaining protein integrity and dynamics. Tightly bound waters stabilize the folded state by bridging hydrogen bonds between protein atoms, contributing to the thermodynamic balance that favors native conformations over unfolded ones. Diffusive waters in the shell enable structural flexibility, facilitating conformational changes essential for enzymatic activity and binding without rigidifying the protein core. Disruption of this shell, such as during thermal or chemical denaturation, leads to loss of these stabilizing interactions, promoting unfolding as access to the hydrophobic core increases. Experimental studies have quantified the hydration shell's composition using techniques like (NMR) and . NMR measurements indicate that functional proteins maintain approximately 0.3-0.5 g of per gram of protein in the shell, corresponding to a coverage that activates dynamics. of globular proteins typically reveals 200-300 ordered molecules in the first hydration shell, often located at polar interfaces and contributing to structural resolution. Evolutionarily, certain hydration shell waters are conserved across homologous proteins, underscoring their structural importance. These conserved waters, identified at ligand-binding interfaces, maintain consistent positions and interactions despite sequence variations, supporting functional specificity. Cosolvents like disrupt shell integrity by displacing water molecules from the first solvation layer, enhancing direct protein- contacts and accelerating denaturation.

Nucleic Acids and Membranes

In nucleic acids, the phosphate backbone is highly hydrated due to its negative charge, with water molecules forming structured layers that stabilize the double helix. Each phosphate group in double-helical DNA coordinates approximately six hydration sites, where water occupancy varies based on the local conformation and sequence. In A-form DNA, a distinctive "spine of hydration" emerges in the minor groove, consisting of ordered water molecules hydrogen-bonded in a zigzag pattern along the grooves, which contributes to the stability of this compact structure under low-humidity conditions. This spine involves about two water molecules per base pair and represents a key feature of the first solvation shell, distinguishing A-form from B-form DNA where hydration is more diffuse. Solvation in the major and minor grooves of nucleic acids plays a critical role in molecular recognition, as water networks mediate interactions with ligands and proteins. Specific patterns of water molecules in the minor groove, such as those forming hydrogen-bonded bridges between bases, determine binding specificity for drugs like , where displacement of groove waters enhances affinity. In AT-rich sequences, narrower minor grooves support denser hydration shells that influence sequence-specific recognition by transcription factors. Counterion effects significantly modulate the solvation shell around nucleic acids, particularly through condensation of cations near the backbone. According to Manning's counterion condensation theory, for DNA with a charge spacing parameter ξ ≈ 4.2 (exceeding 1 for monovalent ions), a layer of condensed Na⁺ ions forms within the first solvation shell, reducing the effective charge and stabilizing the structure against electrostatic repulsion. This condensed layer, typically 5-10 thick, competes with for coordination, altering the dynamics of the hydration shell and facilitating DNA packaging . In lipid membranes, the solvation shell manifests as an interfacial layer at the headgroup region, extending approximately 10 from the bilayer surface and comprising 2-3 ordered layers of molecules. This layer arises from hydrogen bonding between and polar headgroups like , creating a structured zone that influences and permeability. forces between apposing bilayers repel at short distances (below 20 ), preventing collapse and maintaining separation; these forces are quantified using the osmotic stress method, where solutions dehydrate multilamellar vesicles to measure versus spacing. Seminal osmotic stress experiments on bilayers reveal an of repulsion with a of 2-3 , underscoring the role of confined in bilayer stability. Pathological alterations in shells occur during events, where local of the interfacial layer facilitates hemifusion stalks and pore formation. In processes like entry or synaptic , proteins such as SNAREs induce transient removal of 1-2 layers, overcoming the repulsive barrier to merge bilayers. Similarly, aggregates interacting with membranes disrupt the headgroup shell, promoting extraction and pore formation that exacerbates cellular toxicity in diseases like Alzheimer's. These changes highlight how perturbed contributes to destabilization in pathological contexts.

Experimental and Computational Methods

Spectroscopic Techniques

(NMR) spectroscopy provides insights into the dynamics and environment of molecules in the solvation shell through measurements of relaxation times and chemical shifts. The longitudinal (T1) and transverse (T2) relaxation times of protons in shell are typically shorter than those in (~3 s at ) due to restricted motion and interactions with the solute, allowing differentiation of hydration layers. For instance, in protein solutions, effective T2 values for hydration are on the order of 0.1–1 s, reflecting slower rotational correlation times (nanoseconds) compared to 's picosecond regime. perturbations in NMR spectra arise from the altered magnetic environment in the solvation shell, where protons experience deshielding or shielding effects from nearby solute atoms, often shifting by several parts per million relative to . These perturbations are particularly evident in studies of ion solvation, such as alkali metal ions, where the first-shell signals are averaged due to rapid exchange but indicate distinct coordination environments. Infrared (IR) and Raman spectroscopies probe the vibrational modes of water in the solvation shell, revealing differences in hydrogen bonding compared to bulk water. The O-H stretching vibrations of bound water molecules exhibit red-shifted and broadened peaks around 3200–3400 cm⁻¹, contrasting with the sharper ~3700 cm⁻¹ band of free O-H in bulk water, due to stronger hydrogen bonds in the shell. Raman spectroscopy, in particular, highlights these shifts through enhanced scattering from oriented water dipoles near charged solutes, with peak broadening at ~3400 cm⁻¹ indicating heterogeneous environments in the first hydration layer. For example, in aqueous salt solutions, the low-frequency O-H stretch component intensifies, signaling disrupted tetrahedral ordering in the shell. These techniques are sensitive to temperature and concentration effects, showing how solvation alters collective vibrational coupling. Complementary simulations can validate these spectral assignments by modeling shell-specific frequencies. Extended X-ray absorption fine structure (EXAFS) and X-ray absorption near-edge structure (XANES) spectroscopies determine the local atomic structure of solvation shells around metal ions by analyzing X-ray absorption edges. XANES edge shifts reflect changes in and in the ion's shell, while EXAFS provides radial distribution functions to quantify bond lengths and coordination numbers. For Cu²⁺ in aqueous solutions, these methods reveal a first shell of 4–6 oxygen atoms at distances of ~1.95–2.0 Å, indicative of distorted octahedral or square-planar coordination influenced by Jahn-Teller distortion. The technique excels in distinguishing inner-shell ligands from outer-sphere , with Fourier transforms showing peaks corresponding to Cu-O and Cu-Cu distances in concentrated solutions. Such analyses are crucial for understanding ion pairing and shell stability in electrolyte solutions. Dielectric spectroscopy in the terahertz regime captures the collective dielectric response and reorientational dynamics of water in the solvation shell, which differ from bulk water due to slowed rotational diffusion. In this frequency range (0.1–10 THz), the complex dielectric permittivity reveals Debye-like relaxations for shell water, with relaxation times extended to ~10–20 ps compared to bulk water's ~8 ps, reflecting hindered dipole alignment near solutes. For biomolecular systems, terahertz spectra show reduced absorption in the shell, indicating lower mobility and altered hydrogen-bond networks. This approach quantifies the number of dynamically distinct water layers, often identifying 1–2 shells with progressively recovering bulk-like behavior. These measurements complement vibrational probes by focusing on low-frequency collective modes. Neutron scattering techniques, including (SANS) and inelastic neutron scattering (INS), provide information on the structure and dynamics of solvation shells, particularly sensitive to positions. SANS can reveal shell density around proteins or nanoparticles, often showing enhanced density (~10–15% higher than bulk) in the first layer due to favorable interactions. INS probes vibrational and diffusive motions, distinguishing slowed dynamics in the shell (e.g., reduced coefficients) from bulk. For example, in solutions, neutron studies confirm a denser first shell of ~3 thickness. These methods are especially useful for isotopic contrast variation to isolate solute-solvent interactions in complex systems.

Simulation Approaches

Molecular dynamics (MD) simulations provide detailed insights into the structure and dynamics of solvation shells by evolving atomic trajectories under classical force fields. These simulations commonly employ water models such as TIP3P, a three-site rigid model that accurately reproduces water properties and is widely used for studying and biomolecular . Radial functions (RDFs) derived from MD trajectories quantify the probability of finding solvent molecules at specific distances from the solute, defining the first solvation shell up to the first RDF minimum, typically around 3.5 Å for water around ions or proteins. Residence times, calculated from the of solvent occupancy in the shell, reveal dynamical characteristics; for example, in dilute aqueous NaCl solutions, MD simulations show Na⁺ numbers of approximately 5.5 with shell waters exhibiting exchange influenced by and , though specific residence times range from tens to hundreds of picoseconds depending on the . Monte Carlo (MC) methods complement MD by sampling equilibrium configurations of shells in the , focusing on statistical properties without time evolution. These simulations, often using potentials like for , generate ensembles for solutes at infinite dilution to explore shell configurations and thermodynamic stability. perturbations within MC frameworks compute free energies by gradually scaling interactions, enabling assessment of shell stability; for instance, simulations of hydrophobic and charged derivatives in yield free energies increasing with solute size, highlighting the role of shell reorganization in conformational equilibria. Ab initio MD simulations offer a quantum mechanical treatment of solvation shells, capturing electronic effects absent in classical approaches, particularly for small systems or reactive environments. Pure ab initio MD, using , has been applied to formate ion hydration, revealing RDFs with fewer hydrogen bonds per oxygen than classical predictions and average first-shell hydration numbers aligning with experiments. Hybrid quantum mechanics/ (QM/MM) variants extend this to larger systems like proteins by treating the solute and inner shell quantum mechanically while modeling the outer environment classically; for Mg²⁺ in , QM/MM MD simulations demonstrate a stable octahedral first shell of six waters with picosecond-scale exchange and distortions up to 20° in O-Mg-O angles. Recent advances since 2020 incorporate potentials (MLPs) to accelerate large-scale solvation shell simulations, achieving quantum-level accuracy at classical speeds. These potentials, trained on data, model explicit solvent interactions for chemical processes; for example, MLPs have enabled simulations of Diels-Alder reactions in water/, revealing on barriers reduced by 2.4 kcal/mol due to hydrophobic shell stabilization. strategies minimize training data needs, facilitating studies of and shell structures in concentrated electrolytes. Such predictions can be validated against spectroscopic observables like vibrational frequencies.

References

  1. [1]
    1.14: Introduction to Solutions
    ### Definition and Description of Solvation Shell
  2. [2]
    Characterization of the Coordination and Solvation Dynamics of ...
    The definition of a solvation shell is typically based on a given threshold distance determined from the respective solute–solvent pair distribution function g( ...
  3. [3]
    Convergence behaviour of solvation shells in simulated liquids
    Feb 11, 2021 · A convenient way to analyse solvent structure around a solute is to use solvation shells, whereby solvent position around the solute is ...
  4. [4]
    The solvation shell probed by resonant intermolecular Coulombic ...
    Aug 13, 2024 · Introduction. The formation of a solvation shell around ions and molecules dissolved in a liquid is the basis of many properties of solutions.
  5. [5]
    Structure and Dynamics of the Hydration Shell: Spatially ...
    Dec 11, 2018 · Molecular simulations provide insight into solvation structures and dynamics with unparalleled spatial and temporal resolution.Missing: solvation | Show results with:solvation
  6. [6]
    Water Dynamics in the Hydration Shells of Biomolecules
    Mar 1, 2017 · The hydration shell is usually defined in a perturbative perspective: it includes all those water molecules whose properties (e.g., their ...<|control11|><|separator|>
  7. [7]
    Unraveling the Role of Solvation Shell, Surface Adsorption, and SEI ...
    Apr 1, 2025 · Overall, while cosolvents and additives modify the solvation shell, they also protect the anode either through passive adsorption or through ...
  8. [8]
    Solvation - an overview | ScienceDirect Topics
    Solvation is defined as the interaction of a solvent with dissolved molecules, where solvent particles surround and disperse solute particles, influenced by ...
  9. [9]
    Electrolyte Solvation Structure Design for Sodium Ion Batteries - 2022
    Jun 5, 2022 · Compared with the first solvation shell, the secondary solvation shell is less compact; it includes the partially restrained solvent molecules, ...
  10. [10]
    A Theory of Water and Ionic Solution, with Particular Reference to ...
    J. D. Bernal, R. H. Fowler; A Theory of Water and Ionic Solution, with ... This content is only available via PDF. Open the PDF for in another window.
  11. [11]
    Hydrated Sodium Ion Clusters [Na+(H2O)n (n = 1–6)]: An ab initio ...
    Sep 12, 2019 · In liquid water, the coordination number of a sodium ion is about 5.5 ± 0.5 based on molecular dynamics simulation (Mancinelli et al., 2007; ...
  12. [12]
    Impact of hierarchical water dipole orderings on the dynamics of ...
    Aug 7, 2023 · Here we explain these ion-specific properties by the ion-induced hierarchical dipolar, translational, and bond-orientational orderings of ion hydration shell.
  13. [13]
    Solvent Structure, Dynamics, and Ion Mobility in Aqueous Solutions ...
    The solvation dynamics of individual cations and anions are distinct at short times but characterized by the solvent at long times.
  14. [14]
    Hydrogen-bond structure dynamics in bulk water - RSC Publishing
    Dec 4, 2017 · The hydrogen-bond switching dynamics in liquid water are very fast, with a hydrogen-bond life time of around 0.78 picoseconds, determined using AIMD simulation.
  15. [15]
    (PDF) Molecular Dynamic Simulation of the Structure of Solvation ...
    Aug 7, 2025 · The solvation shells of the ions are constructed by analyzing the radial distribution functions and the coordination numbers of the ions.
  16. [16]
    Diffusion coefficient of ionic solvation shell molecules - AIP Publishing
    Feb 24, 2005 · Generally, their motion is found to be substantially slowed down with respect to bulk solvent molecules but still somewhat faster than that of ...
  17. [17]
    Quantum Mechanical Continuum Solvation Models - ACS Publications
    This review on continuum solvation models has been preceded in Chemical Reviews by others addressing the same subject. They are due to Tomasi and Persico 1 ...<|control11|><|separator|>
  18. [18]
    Volumen und Hydratationswärme der Ionen | Zeitschrift für Physik A ...
    Volumen und Hydratationswärme der Ionen. Published: February 1920. Volume 1, pages 45–48, (1920); Cite this article. Download PDF · Zeitschrift für Physik.
  19. [19]
    Electric Moments of Molecules in Liquids - ACS Publications
    Electric Moments of Molecules in Liquids ... Lars Onsager. ACS Legacy Archive. Open PDF. Journal of the American Chemical Society. Cite this: J. Am. Chem. Soc.
  20. [20]
    [PDF] Solute/Solvent Interactions (Theoretical and Computational ...
    The. Born-Kirkwood-Onsager model would require the electrostatic portion of the free energy of solvation for these molecules to be identically zero.<|control11|><|separator|>
  21. [21]
    Ion Solvation in Supercritical Water Based on an Adsorption Analogy
    The model utilizes two parameters: a maximum local density, which is found to be insensitive to temperature and correlated with the surface field, and a ...
  22. [22]
    [PDF] A Simple Model for Ion Solvation with Non Additive Cores - DTIC
    the pair correlation functions can be derived from a simple model in which the ions are replaced by charged shells and the dipoles by charged dipolar shells.
  23. [23]
    Preferential solvation in binary mixtures. Comparison between the ...
    The quasi-lattice quasi-chemical (QLQC) model and the stepwise solvent exchange (SSE) model are both used in solvation chemistry to describe preferential ...Missing: theory | Show results with:theory
  24. [24]
    Significant structure theory applied to electrolyte solution - PubMed
    The significant structure theory has been successfully applied to an aqueous NaCl solution. Liquid water is assumed to have a domain structure.Missing: solvation | Show results with:solvation
  25. [25]
    Solvation Structure and Energetics of Single Ions at the Aqueous ...
    For this analysis, we define the first solvation shell cutoff distance as the position of the first minimum in the radial distribution function for single ions ...
  26. [26]
    Water-Mediated Ion Pairing: Occurrence and Relevance
    We discuss the importance of specific ion-pairing interactions between cations like Li+ and Na+ and anionic carboxylate and phosphate groups for the structure ...
  27. [27]
    A shell-resolved analysis of preferential solvation of coffee ...
    Feb 7, 2018 · Schematic view of solvation shells around an anisotropic solute. The white polyhedra in the dark red region correspond to the Voronoi shell ...
  28. [28]
    Corrected Debye−Hückel Theory of Salt Solutions:  Size Asymmetry and Effective Diameters
    ### Summary of Solvation Shells and Short-Range Interactions in Corrected Debye-Hückel Theory
  29. [29]
    An analysis of the parameters in the Debye-Hückel theory
    Three different activity coefficient models are analyzed and compared for 14 salt systems in order to have a more complete picture of the capabilities and ...
  30. [30]
    The Hydration Number of Li+ in Liquid Water - ACS Publications
    Neutron scattering measurements on LiCl solutions in liquid water have led to a firm conclusion that the Li + ion has six near-neighbor water molecule partners.
  31. [31]
    Solvation Number - an overview | ScienceDirect Topics
    Solvation number represents the number of molecules that remain associated with the ion during its movement through the solution. The solvation number is more ...
  32. [32]
    Ion Pairing | Chemical Reviews - ACS Publications
    Ion pairing describes the (partial) association of oppositely charged ions in electrolyte solutions to form distinct chemical species called ion pairs.
  33. [33]
    Hofmeister Series of Ions: A Simple Theory of a Not So Simple Reality
    Dec 19, 2013 · Ion-specific effects on salting-in and salting-out of proteins, protein denaturation, as well as enzymic activity are typically rationalized in ...
  34. [34]
    Solvation of Electrolytes and Nonelectrolytes in Aqueous Solutions
    ### Summary of Solvation of Nonelectrolytes in Aqueous Solutions
  35. [35]
    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.
  36. [36]
    Water's Structure around Hydrophobic Solutes and the Iceberg Model
    Jan 29, 2013 · The structure of water in the hydration shells of small hydrophobic solutes was investigated through molecular dynamics.
  37. [37]
    A Study on Kirkwood−Buff Integrals and Preferential Solvation in ...
    Analysis of the pressure effect on the local composition in a water-alkanol mixture using Kirkwood-Buff integrals. Journal of Structural Chemistry 2014, 55 ...
  38. [38]
    Protein Hydration Dynamics: Much Ado about Nothing?
    Oct 5, 2017 · (3, 6) Because proteins are large molecules, they can contain more than 1000 water molecules within a thin layer of ∼5Å around its folded state.
  39. [39]
  40. [40]
    Water Determines the Structure and Dynamics of Proteins
    High-resolution experimental techniques, like X-ray crystallography, can determine the position and sometimes orientation of strongly bound water molecules. In ...
  41. [41]
    Water polygons in high-resolution protein crystal structures - NIH
    Average number of waters per molecular weight of protein is 11.8 kDa−1 with a standard deviation of 3.9 kDa−1. This shows that on an average of 12 water ...
  42. [42]
    Hydration in Drug Design. 3. Conserved Water Molecules ... - PubMed
    If water molecules are conserved at the ligand-binding sites of homologous proteins, such a finding may suggest the structural importance of water molecules in ...
  43. [43]
    Urea denaturation by stronger dispersion interactions with proteins ...
    Nov 4, 2008 · In the first step, urea displaces the water molecules within the first solvation shell of the protein that allows it to bind tighter with ...Missing: source | Show results with:source
  44. [44]
    Hydration of the Phosphate Group in Double-Helical DNA
    It is shown that the waters are concentrated in six hydration sites per phosphate and that the positions and occupancies of these sites are dependent on the ...
  45. [45]
    DNA's Chiral Spine of Hydration | ACS Central Science
    May 24, 2017 · This chiral spine of hydration is templated by the helical structure and provides a direct label-free probe of the solvation shell.
  46. [46]
    Chiral SFG spectroscopy detects DNA first hydration shell and base ...
    Sep 4, 2024 · This “spine of hydration” comprises ∼2 water molecules per base pair3 of double-stranded DNA (dsDNA) and accounts for around 3%–6% of the first ...
  47. [47]
    A Role for Water Molecules in DNA−Ligand Minor Groove ...
    Targeting the minor groove of DNA through binding to a small molecule has long been considered an important molecular-recognition strategy in biology.Missing: affecting | Show results with:affecting
  48. [48]
    Structured Water Layers Adjacent to Biological Membranes - PMC
    Two regions of the water correspond to a first and second solvation shell surrounding the phosphocholine groups. The extent of lipid hydration is also revealed ...Missing: stress | Show results with:stress
  49. [49]
    Direct Imaging of Individual Intrinsic Hydration Layers on Lipid ...
    The osmotic stress measurements by McIntosh and Simon have suggested that the hydration force dominates the short-range interaction force if the bilayer ...
  50. [50]
    SNARE-mediated membrane fusion trajectories derived from force ...
    Fusion of lipid bilayers is usually prevented by large energy barriers arising from removal of the hydration shell, formation of highly curved structures, ...
  51. [51]
    Amyloid-Mediated Mechanisms of Membrane Disruption - MDPI
    Many reports suggest that the toxic properties of amyloid aggregates are correlated with their ability to damage cell membranes. However, the molecular ...
  52. [52]
    Assessing the Validity of NMR Relaxation Rates Obtained from ...
    Jun 16, 2023 · The MD simulations of the 1H NMR relaxation times T1,2 in the fast motion regime where T1 = T2 agree with measured (de-oxygenated) T2 data ...
  53. [53]
    Solvent Relaxation NMR Measurements on Polymer, Particle ...
    In this paper the PEO/SDS/silica systems are examined using solvent relaxation nuclear magnetic resonance spectroscopy measurements which can give information ...
  54. [54]
    [PDF] Nuclear Magnetic Resonance Studies of the Solvation Structures of ...
    those in solvation shells, where the exchange frequency is beyond the NMR detection time scale, leading to an averaged peak position for each carbon atoms ...Missing: spectroscopy T2
  55. [55]
    First-Principles Calculation of 1H NMR Chemical Shifts of Complex ...
    Nov 23, 2020 · Probing the Solvent Shell with 195Pt Chemical Shifts: Density Functional Theory Molecular Dynamics Study of PtII and PtIV Anionic Complexes ...
  56. [56]
    Vibrational Coupling and Hydrogen-Bond Structure of Water in the ...
    Aug 2, 2023 · Hydration shell spectroscopy in HOD unambiguously shows that the average H-bonding of water becomes stronger in the HS than that of bulk water.
  57. [57]
    Two-dimensional infrared-Raman spectroscopy as a probe ... - Nature
    Apr 7, 2023 · Finally, we note that a peak around (400 cm−1, 3700 cm−1) appears at elevated temperature, while the other peak at about (700 cm−1, 3400 cm−1) ...
  58. [58]
    EXAFS Study on the Coordination Chemistry of the Solvated Copper ...
    Jul 2, 2020 · The aim of this study is to collect high-quality EXAFS data to high k values of solvated copper(II) ions in solution in series of oxygen donor ...
  59. [59]
    Time-resolved terahertz–Raman spectroscopy reveals that cations ...
    Jun 30, 2022 · The solvation of ions changes the physical, chemical and thermodynamic properties of water, and the microscopic origin of this behaviour is ...
  60. [60]
    Terahertz Spectra and Hydration Dynamics of KNO 3 and K 2 SO 4 ...
    Aug 19, 2025 · On the dynamical side, ultrafast IR spectroscopy has been widely used to track the reorientation and vibrational lifetimes of water molecules in ...
  61. [61]
    Molecular dynamics simulation study of ionic hydration and ion ...
    In the present study, we carried out molecular dynamics simulations of aqueous NaCl solutions at both infinite dilution and rather high concentration (1 molal) ...Original Articles · 3. Infinitely Dilute... · 3.3. Hydration Shell...
  62. [62]
    Hydration and Conformational Equilibria of Simple Hydrophobic and ...
    Canonical ensemble Monte Carlo (MC) simulations (Allen and Tildesley, 1987) of one solute molecule at infinite dilution in water were performed at a ...
  63. [63]
    Ab Initio Molecular Dynamics Study of Formate Ion Hydration
    Ab initio QM/MM approaches generally apply either Hartree−Fock (HF) or density functional theory (DFT) to treat the solute. In virtually all ab initio QM/MM ...
  64. [64]
    Modelling chemical processes in explicit solvents with machine ...
    Jul 20, 2024 · Here, we present a general strategy for generating reactive machine learning potentials to model chemical processes in solution.