Fact-checked by Grok 2 weeks ago

Solubility

Solubility is the analytical composition of a saturated , expressed in terms of the proportion of a designated solute in a designated when the solvent is present in excess. This property quantifies the maximum amount of solute that can dissolve in a given quantity of solvent at under specified conditions, such as and pressure, forming a homogeneous known as a . Solubility is typically expressed in units including grams of solute per 100 grams or milliliters of , molarity (moles per liter), (moles per kilogram of ), or , depending on the context and precision required. A key guiding principle for predicting solubility is "like dissolves like," meaning substances with similar intermolecular forces—such as polar solutes in polar s or nonpolar solutes in nonpolar s—tend to form solutions more readily due to favorable interactions between solute and particles. For ionic compounds, solubility often follows specific rules based on the ions involved, such as nitrates being generally soluble in . Several factors influence solubility, including , , and the chemical nature of the solute and . For most solutes in , solubility increases with rising as the enhances the disruption of solute structures. In contrast, the solubility of gases in liquids generally decreases with increasing , as higher favors the escape of gas molecules from the . significantly affects gas solubility, with higher pressures increasing the amount of gas that dissolves according to , which states that the solubility of a gas is directly proportional to the of the gas above the . These factors are critical in applications ranging from pharmaceutical formulations, where controlled solubility ensures , to environmental processes like ocean carbon cycling influenced by gas solubility in .

Fundamentals of Solubility

Definition and Basic Principles

Solubility refers to the maximum amount of a solute that can dissolve in a given quantity of solvent under specified conditions of temperature, pressure, and composition, resulting in the formation of a saturated solution where the solution is in dynamic equilibrium with any undissolved solute. This property is fundamental to understanding solution formation and is expressed quantitatively as the analytical composition of the saturated solution, often in terms of concentration, mass per volume, or mole fraction. The process of dissolution represents a physical equilibrium, in which the rate of solute particles entering the solution equals the rate at which they crystallize out, maintaining a constant solute concentration once saturation is reached./Equilibria/Solubilty/Solubility_and_Factors_Affecting_Solubility) The concept of solubility has evolved through historical observations and formal definitions. Early insights into gas solubility were provided by English chemist William Henry in the early , who through systematic experiments demonstrated the proportional relationship between gas pressure and its solubility in liquids, laying groundwork for later quantitative studies. Over time, these empirical foundations led to modern standardization by the International Union of Pure and Applied Chemistry (IUPAC), which defines solubility precisely to account for diverse solute-solvent systems and conditions, ensuring consistency in scientific communication and application. To describe the extent of solubility, qualitative terms are commonly used based on approximate thresholds in grams of solute per 100 mL of solvent at standard conditions: substances are considered soluble if exceeding 1 g/100 mL, sparingly soluble if between 0.1 and 1 g/100 mL, and insoluble if less than 0.1 g/100 mL. For instance, (NaCl) exemplifies high solubility in , dissolving at about 36 g/100 mL at 25°C, enabling its widespread use in aqueous solutions. In contrast, (AgCl) is insoluble, with a solubility of only approximately 0.00019 g/100 mL at 25°C, which underlies its precipitation in qualitative analysis. These qualifiers help predict behavior in chemical processes without requiring exact measurements.

Molecular View of Dissolution

The dissolution of a solute in a occurs at the molecular level through the process of , where solute particles are surrounded and stabilized by molecules, forming new intermolecular interactions that replace the original solute-solute attractions. For ionic solutes in polar s like , this involves strong ion-dipole forces, in which the partial charges on the molecules align with the full charges on the ions, effectively pulling them apart from the solid lattice. In cases of non-ionic solutes, weaker van der Waals forces or dipole-dipole interactions may contribute to , facilitating the integration of the solute into the structure. A key aspect of dissolution is the energy balance between breaking solute-solute bonds and forming solute-solvent bonds; for ionic solids, this pits the —the energy required to separate ions in the crystal—against the released when ions interact with water molecules. If the sufficiently offsets the , the process becomes energetically favorable, though the overall spontaneity depends on the change, given by \Delta G = \Delta H - T \Delta S where \Delta H represents the change (primarily from net energy shifts in bond breaking and forming), \Delta S is the change (reflecting increased disorder from dispersing solute particles), and T is the absolute temperature. Negative \Delta G indicates a spontaneous dissolution, with often playing a in overcoming any endothermic contributions. In the , or , molecules form a structured layer around the solute, dynamically orienting to maximize favorable interactions; for example, molecules arrange with their oxygen atoms facing Na^+ ions and atoms toward Cl^- ions, creating a sphere that stabilizes the ions in . This formation enhances solubility by isolating solute particles and preventing recombination, while the 's is essential, as it enables the "like dissolves like" where solutes with similar intermolecular forces to the are more readily solvated.

Quantification of Solubility

Measures per Solvent or Solution

Solubility is often quantified relative to the amount of solvent used, providing a measure independent of the total solution volume. One common expression is grams of solute per 100 grams of solvent (g/100 g), which directly indicates the and is particularly useful for comparing solubilities across different solvents without considering variations. Another key is molality (m), defined as the number of moles of solute per of solvent, which normalizes concentrations based on solvent mass and remains constant despite changes in solution volume. For instance, the solubility of (KNO₃) in at 25°C is approximately 38 g per 100 g of , highlighting its high solubility in this metric. In contrast, measures expressed per quantity of solution account for the total mixture and are suited for volumetric analyses or reactions in fixed volumes. Grams of solute per 100 milliliters of solution (g/100 mL) is a practical unit for laboratory preparations, as it aligns with common measurement tools like pipettes and reflects the density of the solution implicitly. Molarity (M), or moles of solute per liter of solution, facilitates stoichiometric calculations in chemical reactions, while mole fraction (X), the ratio of moles of solute to total moles in the solution, offers a dimensionless scale ideal for thermodynamic discussions. The same KNO₃ solubility at 25°C corresponds to approximately 35 g/100 mL of solution, though exact values depend on whether expressed per solvent mass or solution volume. This approximation holds better for dilute systems where solution density ≈1 g/mL. Each unit has specific advantages depending on the context; for example, is preferred for like because it is unaffected by temperature-induced volume changes, unlike molarity, which varies with density. , meanwhile, simplifies applications by being independent of the solvent's mass or volume. For ionic compounds, is conceptually bounded by the (Ksp), an that equals the product of concentrations (in moles per liter) raised to their stoichiometric powers in a saturated . This value serves as a limit indicating the maximum solubility under ideal conditions, with lower Ksp values signifying poorer solubility, though it applies primarily to sparingly soluble salts and assumes no common effects or formations.

Handling Liquid and Gaseous Solutes

Liquid solutes are handled differently in solubility quantification due to their fluid nature, which allows for complete miscibility or partial mixing without the need to disrupt a rigid structure. Miscibility refers to the ability of two or more liquids to form a homogeneous solution in all proportions, as seen in the ethanol-water system where polar interactions enable full dissolution. In contrast, limited solubility occurs when liquids do not mix completely, such as oil and water, where nonpolar hydrocarbons separate into distinct layers due to unfavorable interactions. Common units for expressing the solubility of liquid solutes include volume percent (v/v), defined as the volume of solute per 100 volumes of solution, and mole fraction, which represents the ratio of moles of solute to total moles in the mixture. Gaseous solutes are quantified primarily through their equilibrium distribution between the gas phase and the liquid solvent, often expressed in volume/volume (V/V) terms or as a function of partial pressure. Henry's law provides the fundamental relationship, stating that at constant temperature, the solubility S of a gas in a liquid is directly proportional to the partial pressure P of the gas above the liquid: S = k_H \cdot P where k_H is the Henry's law constant, which varies with temperature and the gas-solvent pair. This constant can be expressed in units such as atm/mol or mol/(L·atm), allowing conversion between concentration and pressure. A practical example is the dissolution of carbon dioxide (CO₂) in carbonated beverages, where high pressure during bottling increases k_H \cdot P, dissolving CO₂ to create fizz upon opening as pressure drops and the gas escapes. Unlike solid solutes, which require energy to overcome for , liquid and gaseous solutes lack a crystalline , so their solubility emphasizes partitioning driven by intermolecular forces and phase distribution. This partitioning reflects a dynamic balance where the solute distributes between phases based on solubility parameters, without the enthalpic barrier of lattice disruption characteristic of solids.

Unit Conversions and Equivalents

Converting between different units of solubility is essential for comparing data across , applications, and regulatory standards, as solubility is often reported in context-specific formats such as per , moles per , or s. Common conversions involve transforming (moles of solute per of ) to molarity (moles of solute per liter of ) or mass-based units like grams of solute per 100 grams of to . These transformations require knowledge of molar masses, densities, and sometimes approximations for dilute solutions. The conversion from molality to molarity accounts for the volume of the solution, which depends on the solvent's density and the solute's contribution to the total mass. For dilute aqueous solutions, an approximation is M ≈ m (since 1 kg water ≈ 1 L). For more accurate results, especially in concentrated solutions, the full calculation uses the actual solution density (\rho): M = \frac{m \cdot \rho}{1 + m \cdot \frac{M_{\text{solute}}}{1000}} where M is molarity in mol/L, m is molality in mol/kg, \rho is solution density in g/mL, and M_{\text{solute}} is the molar mass of the solute in g/mol. This incorporates the total mass and volume of the solution. Similarly, converting from grams of solute per 100 grams of (g/100 g) to (X) involves calculating the moles of each component. The of the solute is given by: X = \frac{n_{\text{solute}}}{n_{\text{solute}} + n_{\text{solvent}}} where n_{\text{solute}} and n_{\text{solvent}} are the moles of solute and , respectively, determined from their masses and masses. For example, if 36 g of solute is dissolved in 100 g of , divide each mass by the appropriate to find the moles, then apply the formula. A practical example is the solubility of sodium chloride (NaCl) in water at 25°C, reported as 36 g NaCl per 100 g water. First, convert to molality: moles of NaCl = 36 g / 58.44 g/mol ≈ 0.616 mol; molality m = 0.616 mol / 0.100 kg = 6.16 m. To find molarity, note the total mass of solution = 136 g and density of the saturated solution ≈ 1.202 g/mL, so volume ≈ 136 g / 1.202 g/mL ≈ 0.113 L; thus, M ≈ 0.616 mol / 0.113 L ≈ 5.4 M. This conversion highlights the need for solution density, which is often assumed as 1 g/mL for dilute cases but leads to errors in concentrated solutions like this one. In different systems, particularly for trace solubilities, parts per million (ppm) is widely used, defined as milligrams of solute per kilogram of solution (or equivalently, \mug/g for mass basis). This unit is common in environmental contexts, such as assessing pollutant solubility in water bodies, where concentrations below 1 mg/L (≈ 1 ppm assuming density ≈ 1 g/mL) indicate trace levels. For instance, the solubility of sparingly soluble salts like lead(II) sulfate in natural waters is often expressed in ppm to evaluate environmental risks. Errors in unit conversions can arise from variations in and density, which affect both the reported solubility and the conversion factors. Solubility itself often changes with (e.g., increasing for most in ), and decreases as rises, potentially introducing up to 5-10% error in molarity calculations if not adjusted. For precise work, use temperature-specific values and re-evaluate solubility data at the exact conditions.

Factors Influencing Solubility

Temperature Dependence

The solubility of solutes in solvents generally varies with , depending on whether the dissolution process is endothermic or exothermic. In endothermic dissolution, where the process absorbs , solubility increases as rises, shifting the toward greater to absorb the added . For instance, the solubility of (KNO₃) in exemplifies this trend, increasing from about 13 g per 100 mL at 0°C to 247 g per 100 mL at 100°C. Conversely, exothermic dissolution releases , leading to decreased solubility with increasing as the system shifts to counteract the added by favoring the undissolved state. (Ca(OH)₂) in demonstrates this behavior, with solubility dropping from 0.173 g per 100 mL at 20°C to 0.066 g per 100 mL at 100°C. This temperature dependence can be understood through Le Chatelier's principle, which predicts that the equilibrium position will adjust to minimize changes in conditions. For dissolution, the process is treated as a reversible equilibrium: solute(s) ⇌ solute(aq) + heat (for exothermic) or solute(s) + heat ⇌ solute(aq) (for endothermic). Increasing temperature thus drives endothermic equilibria forward, enhancing solubility, while opposing exothermic ones, reducing it./Equilibria/Solubilty/Temperature_Effects_on_Solubility) Quantitatively, the relationship between temperature and solubility is described by the van't Hoff equation, derived from the temperature dependence of the equilibrium constant K (here, related to the solubility product or saturation concentration). The equation is: \frac{d(\ln K)}{dT} = \frac{\Delta H}{RT^2} where \Delta H is the enthalpy change of dissolution, R is the gas constant, and T is the absolute temperature. Integrating this form allows prediction of how solubility varies with temperature, assuming \Delta H is constant; a positive \Delta H (endothermic) yields increasing K with T, while negative \Delta H (exothermic) yields decreasing K./26%3A_Chemical_Equilibrium/26.07%3A_The_van_%27t_Hoff_Equation) Exceptions to these patterns occur, particularly with hydrated salts undergoing phase transitions. For example, decahydrate (Na₂SO₄·10H₂O) exhibits retrograde solubility above approximately 32°C, where its solubility peaks and then slightly decreases due to and transition to the form, altering the effective .

Pressure Effects

The solubility of gases in s is significantly influenced by , as described by , which states that at constant temperature, the solubility S of a gas in a is directly proportional to the partial P of the gas above the :
S = k \cdot P
where k is the Henry's law constant, typically expressed in units such as mol/L/atm or mol/L/bar. This relationship arises from the increased frequency of gas collisions with the surface under higher , leading to greater until is reached./Physical_Properties_of_Matter/Solutions_and_Mixtures/Ideal_Solutions/Dissolving_Gases_In_Liquids_Henrys_Law)
For instance, the solubility of oxygen in human , which follows , approximately doubles when the total pressure increases from 1 to 2 , assuming the partial pressure of oxygen also doubles in ; this enhances oxygen delivery but is limited by blood's low baseline solubility of about 0.003 mL O₂/100 mL blood/mm Hg at 37°C. constants for common gases vary widely; for oxygen in at 25°C, k is approximately 0.0013 /L/, while for it is higher at 0.034 /L/, reflecting differences in gas-liquid interactions. In contrast, the solubility of liquids in liquids and solids in liquids exhibits minimal dependence on under normal conditions, due to the low of these , which results in negligible changes in during dissolution./13:_Solutions/13.04:_Effects_of_Temperature_and_Pressure_on_Solubility) However, under extreme pressures, such as those encountered in deep-sea environments (up to several hundred ) or industrial high-pressure processes, solubility of solids can increase slightly if the partial molar volume of the solute in is less than in the phase, as predicted by thermodynamic relations. A practical application of pressure effects on gas solubility is in , where increased ambient at depth raises the solubility of in blood and tissues according to ; upon rapid ascent and pressure reduction, dissolved forms bubbles, potentially causing if not managed through staged decompression stops.

Polarity and Intermolecular Forces

The solubility of a substance is fundamentally governed by the principle that "like dissolves like," where polar solutes tend to dissolve in polar solvents, and nonpolar solutes dissolve in nonpolar solvents, due to compatible intermolecular attractions. For instance, glucose, a polar with multiple hydroxyl groups, dissolves readily in through hydrogen bonding between its -OH groups and . In contrast, nonpolar dissolves in nonpolar primarily via dispersion forces between their structures. This compatibility arises because the solute-solvent interactions must overcome the solute-solute and solvent-solvent forces for dissolution to occur effectively. Intermolecular forces play a hierarchical role in determining solubility, ranked by strength as follows: ion-ion interactions are the strongest, followed by ion-, - (including hydrogen bonding as a specialized form), -induced , and the weakest dispersion forces. In polar solvents like , stronger forces such as ion- and hydrogen bonding stabilize ionic or polar solutes by surrounding them with oriented solvent molecules, enhancing solubility. Nonpolar solutes, lacking permanent s, rely on weaker dispersion forces, making them more soluble in nonpolar environments where similar weak attractions predominate. The dielectric constant (ε) of a quantifies its and ability to screen charges, directly influencing solubility; high ε values, such as water's ε ≈ 80 at 25°C, promote the dissolution of ionic compounds by reducing electrostatic attractions between ions. Conversely, low ε like (ε ≈ 2) exhibit poor of ions but effectively dissolve nonpolar molecules through minimal charge separation. This property underscores why ionic salts are highly soluble in but insoluble in hydrocarbons. A classic example of polarity's impact is floating on : nonpolar molecules interact weakly via forces among themselves and cannot form favorable interactions with polar , leading to due to 's strong bonding network. address this incompatibility by featuring both polar (hydrophilic) heads and nonpolar (hydrophobic) tails, enabling them to bridge the two phases, stabilize emulsions, and enhance solubility of nonpolar substances in aqueous media. energy, derived from these molecular interactions, briefly quantifies the net stabilization achieved during dissolution.

Solubility in Specific Systems

Gases in Liquids

The solubility of gases in liquids is fundamentally described by Henry's law, which posits that at a constant temperature, the solubility of a gas in a liquid is directly proportional to the partial pressure of the gas in equilibrium with the liquid phase. This relationship is expressed as c = k_H \cdot p, where c is the concentration of the dissolved gas, p is the partial pressure, and k_H is the Henry's law constant specific to the gas-solvent pair. The law assumes dilute solutions and ideal behavior, providing a foundational framework for predicting gas dissolution in processes ranging from industrial gas absorption to environmental equilibria. However, Henry's law has limitations, particularly at high pressures where non-ideal gas behavior and solute-solvent interactions cause deviations from linearity. For instance, at elevated pressures, the solubility of gases like CO₂ can exceed predictions due to enhanced molecular clustering or effects, complicating applications in high-pressure systems such as deep-sea environments or supercritical extractions. also profoundly influences gas solubility; unlike most , gases exhibit an inverse temperature dependence because dissolution is typically exothermic, favoring lower temperatures for higher solubility. The Henry's constant generally decreases with temperature for common gases in aqueous solvents, consistent with the overall decline in solubility. A representative example is oxygen in , where solubility drops from about 9.1 mg/L at 20°C to 6.4 mg/L at 40°C under , roughly a 30% reduction that impacts oxygen availability in warming aquatic systems. Additional factors modulate gas solubility in liquids. Electrolytes induce a salting-out effect by structuring molecules around ions, reducing the availability of sites for gas molecules and thereby decreasing solubility. This is evident in CO₂, which shows lower solubility in (salinity ~35 ppt) compared to at the same and , with reductions up to 10-20% attributed to ionic competition. enhances solubility as per , but this section focuses on gas-specific behaviors beyond general effects. Practical examples illustrate these principles. In aquatic ecosystems, the oxygenation of water supports , with dissolved oxygen levels critically dependent on and ; warmer waters hold less oxygen, stressing and in climate-altered habitats. Similarly, carbonation of beverages relies on pressurizing CO₂ into liquids at low temperatures to achieve , releasing bubbles upon for . In a modern environmental context, oceans absorb approximately 23-30% of CO₂ emissions, buffering atmospheric levels but driving . Since the 1980s, surface ocean has declined by about 0.1 units (a 30% acidity increase), with 2020s data indicating sustained uptake rates of ~9-10 billion metric tons of CO₂ annually despite warming trends that reduce solubility per . This has led to undersaturation of carbonate minerals, threatening calcifying organisms like corals and , as documented in recent global assessments.

Ionic Compounds in Water

The solubility of ionic compounds in arises from the balance between the energy required to disrupt the ionic lattice and the energy released through hydration. The represents the strong electrostatic attractions holding ions together in the solid crystal, which must be overcome for to occur. In contrast, is the exothermic process where molecules surround and stabilize the separated ions via ion-dipole interactions. Ionic compounds are generally highly soluble if the magnitude of the exceeds the , resulting in a net favorable change for . For example, (NaCl) exhibits high solubility of approximately 360 g/L at 25°C, as its sufficiently compensates for the . Empirical solubility rules provide guidelines for predicting the behavior of many ionic compounds in , based on observed patterns. All nitrates (NO₃⁻) are soluble, regardless of the cation, due to the weak forces in these salts. Carbonates (CO₃²⁻), however, are typically insoluble except when paired with cations (e.g., Na⁺, K⁺) or (NH₄⁺), where the high of these small, highly charged cations promotes . These rules stem from the interplay of sizes, charges, and hydration tendencies, allowing chemists to anticipate or formation without exhaustive experimentation. The further modulates solubility by introducing an ion already present in the equilibrium, shifting the dissolution process toward the solid phase per . For instance, the solubility of (AgCl), which is sparingly soluble in pure , decreases significantly in a containing chloride ions from (NaCl). The additional Cl⁻ ions suppress AgCl , reducing its molar solubility from about 1.3 × 10⁻⁵ M in pure to ≈1.8 × 10^{-9} M in 0.10 M NaCl. This effect is crucial in qualitative analysis and precipitation reactions. Certain ionic hydroxides display amphoteric behavior, dissolving in both acidic and basic conditions due to pH-dependent . Aluminum hydroxide (Al(OH)₃) exemplifies this, with minimal solubility near neutral (around 6–8), where it precipitates as the neutral hydroxide. In acidic media ( < 5), it dissolves by forming soluble aluminate ions or complexes like [Al(H₂O)₆]³⁺, while in basic media ( > 9), it forms tetrahydroxoaluminate ions ([Al(OH)₄]⁻). This pH-dependent solubility curve reflects the compound's ability to act as either an or , enhancing its utility in buffering and processes.

Organic Compounds in Solvents

The solubility of compounds in solvents is largely governed by the of the solute and solvent, with polar functional groups facilitating dissolution in polar media like through bonding and interactions. Functional groups play a critical role in determining solubility patterns; for instance, the presence of a hydroxyl (-OH) group in alcohols enhances solubility by enabling bonding with molecules, as seen in short-chain alcohols like . In contrast, moieties, being nonpolar, reduce solubility in by limiting such interactions, making purely -based compounds like alkanes nearly insoluble. Other polar groups, such as amines or carboxylates, similarly promote solubility in aqueous solvents, while nonpolar groups like alkyl chains diminish it. A key quantitative measure of an organic compound's hydrophobicity is the logarithm of the (LogP), which indicates the distribution between a nonpolar (octanol) and polar () phase; values greater than 3 typically signify low solubility and high , aiding predictions for pharmaceutical and environmental applications. This metric correlates inversely with aqueous solubility, as higher LogP values reflect stronger partitioning into nonpolar environments. Solubility trends in homologous series, such as alcohols, show a decrease with increasing chain length due to the growing dominance of hydrophobic alkyl portions over the polar hydroxyl group; methanol and ethanol are fully miscible in water, but solubility drops sharply for longer chains, with hexanol exhibiting limited solubility of about 5.9 g/100 mL at 20°C. This pattern underscores the balance between polar and nonpolar contributions in molecular design. To enhance solubility of poorly water-soluble organics, particularly drugs, cosolvents like are commonly added to aqueous systems, reducing the overall and improving through mechanisms such as weakened water structure and increased solute-solvent interactions; for example, mixtures can increase ibuprofen solubility by up to 10-fold in pharmaceutical formulations. This cosolvency approach is vital in , enabling higher without altering the active molecule. Recent advancements emphasize bio-based green solvents as sustainable alternatives for dissolving organic compounds, addressing environmental concerns with traditional solvents; post-2020 developments include the use of bio-derived systems like or cyrene for extracting lipophilic organics such as , offering comparable solubility enhancements while being biodegradable and low-toxicity. These solvents, often sourced from renewable feedstocks, support greener processes in industries like pharmaceuticals and fine chemicals.

Solid Solutions and Alloys

Solid solutions in alloys represent a form of solubility where one metal (the solute) dissolves into the crystal lattice of another metal (the ) to form a homogeneous crystalline without . This occurs at the atomic level, enhancing material properties such as strength and . Unlike liquid solutions, solid solubility is limited by thermodynamic factors and typically requires compatible atomic structures. There are two primary types: substitutional and solid solutions. In substitutional solid solutions, solute atoms replace solvent atoms in the sites, requiring similar atomic sizes and crystal structures for . A classic example is the - (Cu-Ni) alloy, where atoms substitute for in the face-centered cubic , forming a complete solid solution across all compositions at elevated temperatures. Interstitial solid solutions, by contrast, involve small solute atoms occupying the voids (interstices) between larger solvent atoms, without displacing them. Carbon in iron exemplifies this, where carbon atoms fit into the octahedral sites of the body-centered cubic iron in low-carbon steels, enabling limited solubility up to about 0.02 wt% at . These mechanisms allow alloys to achieve uniform properties but are constrained by the solute's size relative to the host . The extent of solid solubility is governed by the , empirical guidelines established in the 1930s that predict conditions for extensive substitutional solubility. These include: (1) a relative atomic size difference of less than 15% between solute and solvent; (2) similar crystal structures; (3) comparable electronegativities to ensure favorable bonding; and (4) the same count for electronic compatibility. For interstitial solubility, the solute must be significantly smaller (atomic radius ratio <0.59) to fit lattice gaps without excessive strain. These rules explain why elements like dissolve well in silver but not in magnesium, influencing alloy design for specific applications. Phase diagrams for binary alloys illustrate solubility limits through key features like the solvus line and eutectic points. The solvus line demarcates the boundary between a single solid solution phase and a two-phase region of solid solutions plus precipitates, showing how maximum solute solubility decreases with falling temperature in many systems. For instance, in the Cu-Ni diagram, the solvus is absent due to complete solubility, but in Al-Cu, it defines the α-phase solubility limit. Eutectic points mark the lowest melting temperature where a liquid decomposes into two solid phases, often bounding limited solid solubility regions; the lead-tin (Pb-Sn) eutectic at 183°C exemplifies this, with minimal mutual solubility in the solids. Temperature dependence thus plays a critical role, as cooling below the solvus can drive precipitation and alter properties. Applications of solid solutions leverage these principles for enhanced performance. In steel production, interstitial carbon forms solid solutions with iron, enabling where solute atoms distort the lattice and impede dislocation motion, increasing yield strength by up to 50% in low-alloy steels without brittleness. Substitutional alloys like (Fe-Cr-Ni) use Hume-Rothery-compliant elements for resistance via uniform lattice integration. In modern , solid solution perovskites—such as doped ABO₃ structures like La₀.₈Sr₀.₂MnO₃—have advanced technologies by improving ionic conductivity in solid-state electrolytes; recent lattice-matched antiperovskite-perovskite interfaces achieve near-theoretical lithium-ion transport rates, boosting in all-solid-state batteries toward commercialization by 2030. These developments highlight solid solutions' role in storage.

Dissolution Dynamics

Rate of Dissolution

The rate of dissolution describes the kinetics by which a solid solute disperses into a solvent, approaching the equilibrium solubility concentration over time. This dynamic process is governed by mass transfer mechanisms and is distinct from the equilibrium solubility, which represents the maximum solute concentration achievable; for instance, a sparingly soluble compound in fine powder form can exhibit a rapid initial dissolution rate, quickly saturating the solution despite its low ultimate solubility. Several key factors influence the dissolution rate. Increasing the surface area of the solute, such as by using powdered rather than crystalline forms, directly enhances the rate by providing more sites for interaction; for example, micronized aspirin tablets dissolve significantly faster across ranges of 1.2 to 6.8 compared to standard 500 mg tablets due to their greater surface area per unit mass. of the accelerates dissolution by thinning the hydrodynamic adjacent to the solute surface, thereby facilitating solute to the bulk . exerts a profound effect, as higher temperatures increase both the solute's saturation concentration and the molecular coefficient, with the temperature dependence of the rate constant following the k = A e^{-E_a / RT}, where A is the , E_a is the , R is the , and T is the absolute ; in dissolution studies of minerals like , activation energies around 89 kJ/mol have been reported, illustrating the exponential sensitivity to . The foundational mathematical model for dissolution kinetics is the Noyes-Whitney equation, originally derived from experiments on solids dissolving in their own solutions: \frac{dC}{dt} = \frac{D A}{h V} (C_s - C) Here, \frac{dC}{dt} is the rate of change of solute concentration in the bulk solution, D is the diffusion coefficient of the solute, A is the surface area of the exposed solid, h is the thickness of the diffusion boundary layer, V is the volume of the solvent, C_s is the saturation solubility, and C is the current bulk concentration. This equation highlights that dissolution is often diffusion-controlled under typical conditions, with the driving force being the concentration gradient across the boundary layer. The process unfolds in sequential stages: first, of solute molecules or ions from the solid at the ; second, of these detached through the unstirred to the bulk ; and third, , where solvent molecules surround and stabilize the solute particles. In many pharmaceutical applications, such as the disintegration of aspirin tablets in aqueous media, the stage predominates as the rate-limiting step, particularly for poorly water-soluble drugs, emphasizing the importance of strategies to optimize dynamics and surface exposure.

Incongruent Dissolution Processes

Incongruent refers to the process in which a solid solute partially reacts or decomposes during , resulting in a saturated whose composition differs from that of the original solid, accompanied by the formation of a new solid that is typically more stable under the given conditions. Unlike congruent , where the solid fully dissolves into its constituent in stoichiometric proportions, incongruent involves selective of components, leading to secondary precipitates or altered solid residues. This phenomenon is common in multicomponent systems, such as hydrated salts or silicates, where environmental factors like , , or activity drive the toward . This process is influenced by and temperature in systems like hydrated calcium sulfates, where a less soluble phase may form as a residue. in certain ionic compounds can also lead to of hydroxides due to local pH changes. in hydrated salts, where lose water to form less hydrated phases upon exposure to dry air, represents a related without full . The Gibbs phase rule, F = C - P + 2, where F is the , C is the number of components, and P is the number of phases, governs these systems, particularly in condensed solid-liquid equilibria without vapor involvement (effectively F = C - P + 1). In incongruent dissolution, the presence of multiple solid phases (e.g., original solute and precipitate) alongside the liquid often results in invariant or univariant conditions, limiting the system's variability and dictating the equilibrium compositions. For instance, in a like a hydrated salt-water, the coexistence of multiple solid phases and saturated creates an invariant point. In industrial contexts, incongruent dissolution plays a critical role in cement hydration, where (C-S-H) gels, the primary binding phase in , dissolve non-stoichiometrically during or , preferentially releasing calcium while leaving a silica-enriched residue, which affects long-term durability and retention in waste repositories. In chemistry and , incongruent processes drive , such as the partial of feldspars to form secondary clays like , influencing nutrient cycling, , and ; studies on silicate highlight silica retention in soils, impacting global geochemical fluxes.

Theoretical Models of Solubility

Like Dissolves Like Principle

The "like dissolves like" principle is an empirical guideline in chemistry stating that substances with similar intermolecular forces tend to be mutually soluble, while those with dissimilar forces are typically immiscible. This rule arises from observations that polar solutes dissolve preferentially in polar solvents and nonpolar solutes in nonpolar solvents, due to comparable cohesive energies that minimize the energy change upon mixing. The principle originated as qualitative observations by early chemists studying solution behavior, but it was formalized quantitatively through the Hildebrand solubility parameter in the 1930s. Joel H. Hildebrand introduced this parameter, δ, defined as the square root of the cohesive energy density: \delta = \sqrt{\frac{\Delta H_v - RT}{V_m}} where ΔH_v is the molar heat of vaporization, R is the gas constant, T is temperature, and V_m is the molar volume; at room temperature, the RT term is often negligible, simplifying to δ ≈ sqrt(ΔH_v / V_m). Solubility is favored when the δ values of solute and solvent differ by less than about 2 (cal/cm³)^{1/2}, reflecting similar intermolecular attractions. This parameter finds practical applications in predicting the of polymers and organic compounds in . For instance, (PVC), with δ ≈ 9.6 (cal/cm³)^{1/2}, dissolves readily in (THF), which has δ ≈ 9.1 (cal/cm³)^{1/2}, due to their closely matched values that facilitate uniform mixing in coatings and adhesives. Such predictions guide selection in industries like pharmaceuticals and paints, where matching δ values ensures effective without . Despite its utility for nonpolar systems, the principle has limitations in cases involving specific interactions, such as bonding or ionic , where the single δ parameter cannot capture directional or charge-based forces. For example, it inadequately predicts solubility in protic solvents like for solutes with strong bonding, as the parameter overlooks gradients. To address these shortcomings, Charles M. extended the approach in 1967 by decomposing δ into three components: (δ_d) for van der Waals forces, polar (δ_p) for dipole-dipole interactions, and -bonding (δ_h) for donor-acceptor bonds, with the total δ_t = sqrt(δ_d² + δ_p² + δ_h²). This solubility parameter framework improves accuracy for complex systems, such as polymers with mixed interactions, by calculating a multidimensional "" between solute and parameters; solubility occurs when this is below a material-specific .

Solubility Product Constant

The solubility product constant, denoted K_{sp}, is the for the of a sparingly soluble ionic in , quantifying the extent to which the dissociates into its constituent ions at . For a general ionic compound A_m B_n(s) \rightleftharpoons m A^{m+}(aq) + n B^{n-}(aq), the expression is K_{sp} = [A^{m+}]^m [B^{n-}]^n, where concentrations are in moles per liter and activities are approximated by concentrations for dilute solutions. This constant is characteristic of each at a given and serves as a measure of its solubility under equilibrium conditions. To calculate the solubility of a 1:1 electrolyte like silver chloride (\ce{AgCl(s) \rightleftharpoons Ag+(aq) + Cl-(aq)}), where K_{sp} = [\ce{Ag+}] [\ce{Cl-}] = 1.8 \times 10^{-10} at 25°C, let s be the molar solubility; then [\ce{Ag+}] = s and [\ce{Cl-}] = s, yielding K_{sp} = s^2 and s = \sqrt{K_{sp}} = \sqrt{1.8 \times 10^{-10}} \approx 1.3 \times 10^{-5} M. For compounds with different stoichiometries, such as \ce{CaF2(s) \rightleftharpoons Ca^{2+}(aq) + 2F-(aq)}, the expression becomes K_{sp} = [ \ce{Ca^{2+}} ] [ \ce{F-} ]^2 = s (2s)^2 = 4s^3, so s = \sqrt{{grok:render&&&type=render_inline_citation&&&citation_id=3&&&citation_type=wikipedia}}{K_{sp}/4}. The presence of a common ion suppresses solubility, as described by Le Châtelier's principle, shifting the toward the undissolved solid. For instance, in a containing 0.10 M \ce{NaCl}, the solubility of \ce{AgCl} decreases because added \ce{Cl-} increases [\ce{Cl-}], requiring a lower s to maintain K_{sp}; solving $1.8 \times 10^{-10} = s (0.10 + s) \approx s \times 0.10 gives s \approx 1.8 \times 10^{-9} M, far less than in pure . Precipitation occurs when the ion product Q = [A^{m+}]^m [B^{n-}]^n exceeds K_{sp}, indicating a supersaturated solution that will form a solid until equilibrium is reached. If Q < K_{sp}, the solution is unsaturated and no precipitate forms; if Q = K_{sp}, it is saturated. This criterion is used to predict whether mixing solutions will result in insolubility, such as when combining silver nitrate and sodium chloride exceeds K_{sp} for \ce{AgCl}. The value of K_{sp} depends on , following the van't Hoff equation, where endothermic dissolution increases K_{sp} with rising , while exothermic processes decrease it. For (\ce{CaCO3(s) \rightleftharpoons Ca^{2+}(aq) + CO3^{2-}(aq)}), solubility decreases with increasing , with the temperature derivative of solubility ranging from -10^{-6} to -3 \times 10^{-5} molal/°C at constant CO₂ , reflecting its exothermic nature.

Advanced Theories and Extensions

The Debye-Hückel theory addresses deviations from ideality in electrolyte solutions by modeling the electrostatic interactions surrounding each ion as an ionic atmosphere, which screens the central ion's charge and reduces its effective activity. Developed by Peter Debye and Erich Hückel in 1923, this theory applies primarily to dilute solutions where interionic forces dominate over short-range interactions. In the limiting case for very low concentrations, the mean activity coefficient \gamma_\pm for a single charge type is given by: \log \gamma_\pm = -0.51 |z_+ z_-| \sqrt{I} where z_+ and z_- are the ion charges, I is the ionic strength, and the constant 0.51 is specific to aqueous solutions at 25°C. This equation corrects the solubility product constant K_{sp} for non-ideal behavior, enabling more accurate predictions of ionic solubility in low-concentration regimes. Extensions like the Davies equation incorporate higher-order terms for moderate concentrations, but the core theory remains foundational for interpreting ion pairing and activity in sparingly soluble salts. Regular solution , introduced by Joel H. Hildebrand in 1929, extends the concept of solubility to nonpolar and weakly polar compounds by assuming random mixing without volume change or of mixing beyond forces. It posits that solubility arises from the balance between cohesive energies in solute and solvent, quantified through solubility parameters \delta, where is favored when \delta_1 \approx \delta_2. The predicts the \gamma via: RT \ln \gamma = V (\delta_1 - \delta_2)^2 \phi_2^2 with V as the molar volume of the solute, \phi_2 as the volume fraction, and \delta as the Hildebrand solubility parameter derived from cohesive energy density. This framework has been widely applied to predict phase behavior in nonpolar solvents, such as hydrocarbon mixtures, though it underperforms for systems with hydrogen bonding or polar interactions due to neglected specific forces. Hildebrand's work laid the groundwork for later thermodynamic models in chemical engineering. For polymeric systems, the Flory-Huggins theory, independently formulated by Paul J. Flory and Morris L. Huggins in 1941-1942, describes the thermodynamics of polymer-solvent mixtures using a lattice model that accounts for the large size disparity between components. The free energy of mixing \Delta G_m is expressed as: \frac{\Delta G_m}{RT} = n_1 \ln \phi_1 + n_2 \ln \phi_2 + \chi n_1 \phi_1 \phi_2 where n_1 and n_2 are the numbers of solvent and polymer molecules, \phi_1 and \phi_2 are their volume fractions, and \chi is the Flory-Huggins interaction parameter reflecting enthalpic contributions. This model highlights the entropic penalty of polymer chain confinement, predicting lower critical solution temperatures for many polymer blends and guiding solubility in applications like coatings and drug delivery. Limitations arise from its mean-field approximation, which overlooks chain connectivity and concentration gradients. Group contribution methods like , developed by Aa. Fredenslund, R.L. Jones, and J.M. Prausnitz in 1975, enable predictive calculations of activity coefficients for complex organic mixtures by decomposing molecules into functional groups and assigning interaction parameters based on experimental data. The model combines a combinatorial term for size and shape effects with a residual term for group-group energies, allowing estimation of solubility without pure-component data. has been extended to over 100 main groups, improving predictions for vapor-liquid equilibria and liquid-liquid separations in . In the 2020s, and approaches have emerged as powerful extensions for solubility prediction, particularly in , where traditional models struggle with diverse chemical spaces. Graph neural networks and random forests trained on large datasets like ESOL and AQUA achieve root-mean-square errors around 0.6 log units for aqueous solubility, often outperforming group contribution methods for novel compounds by learning implicit molecular features. These models integrate quantum descriptors and experimental assays to accelerate lead optimization, as demonstrated in for pharmaceuticals, reducing reliance on costly syntheses. Hybrid AI-UNIFAC frameworks further enhance predictive power by combining data-driven learning with physicochemical rules. Despite these advances, ideal assumptions in basic theories fail in concentrated solutions, where short-range interactions, ion pairing, and non-Coulombic forces lead to significant deviations; for instance, Debye-Hückel overestimates screening at ionic strengths above 0.1 M, necessitating specific-ion models like for accurate solubility in brines or biological fluids. Regular solution and Flory-Huggins theories similarly break down in associating systems, highlighting the need for multiscale simulations to capture dynamics.

Practical Aspects and Predictions

Key Applications

In the , solubility plays a pivotal role in and delivery, particularly for poorly water-soluble compounds that constitute over 40% of new drug candidates. Amorphous forms of s, which lack the ordered crystal lattice of their crystalline counterparts, exhibit significantly higher solubility due to their higher , enabling enhanced rates and improved . Converting a to its amorphous can theoretically increase solubility by factors up to 10- to 100-fold for various compounds, though practical enhancements are often lower due to recrystallization tendencies. Studies on using solid dispersions have shown increases of 2- to 5-fold, thereby facilitating better in the and achieving therapeutic plasma concentrations. This approach is widely adopted in solid dispersions and spray-dried s to overcome limitations, directly impacting the efficacy of oral medications. Environmental applications of solubility concepts are crucial for remediation and . In remediation, the solubility of such as lead, , and increases under acidic conditions ( below 7), facilitating their from contaminated sites but also enabling targeted extraction through or techniques. For example, bio-chelate assisted enhances the mobilization of metals like and by adjusting and using organic agents to form soluble complexes, allowing for their removal and reducing contamination risks. In processes, solubility governs the of soluble impurities; and coagulation-flocculation methods exploit solubility products to convert dissolved ions like calcium, magnesium, and phosphates into insoluble precipitates, which are then settled and filtered to produce potable . These strategies are essential for treating industrial effluents and municipal , preventing the spread of toxic solubilized contaminants. Differential solubility under varying conditions, such as , underpins separation techniques like fractional crystallization, which is used to purify compounds from mixtures. A classic example is the separation of (KNO₃) from (NaCl) in aqueous solutions: KNO₃ has a highly temperature-dependent solubility—increasing dramatically from about 13 g/100 mL at 0°C to 247 g/100 mL at 100°C—while NaCl's solubility remains relatively constant at around 36 g/100 mL. By dissolving the mixture in hot water and cooling the solution, KNO₃ crystallizes preferentially due to its reduced solubility at lower temperatures, allowing for its isolation with high purity, a method applied in chemical manufacturing and purification. This principle extends to for isolating salts and organics, leveraging solubility gradients to achieve efficient separations without complex equipment. In food and chemical processing, solubility drives extraction techniques for isolating valuable components from raw materials. Solvent extraction relies on the differential solubility of target solutes, such as oils, flavors, or bioactive compounds, in immiscible solvents like or , enabling selective partitioning from complex matrices like seeds or herbs. For instance, in edible oil production, solvents are chosen based on Hansen solubility parameters to maximize the extraction of while minimizing unwanted polar impurities, yielding high-purity products with efficiencies up to 99%. These methods are integral to and quality control, ensuring the removal of contaminants or concentration of nutrients. Additionally, in the context of environmental impacts, —exacerbated by climate-driven warming—reduces the solubility of oxygen in , with projections indicating a 3-4% decline in dissolved oxygen by 2100 under high-emission scenarios, leading to expanded hypoxic zones that threaten ecosystems and fisheries as of 2025 assessments.

Methods for Solubility Prediction

Methods for predicting solubility encompass empirical correlations, theoretical frameworks, and computational models that estimate solubility parameters without direct experimentation. These approaches are essential in fields like pharmaceutical development and , where rapid screening of compounds is required. Empirical methods rely on observed patterns from experimental data, while theoretical and computational techniques incorporate molecular and interactions for broader applicability. Empirical predictions often draw from solubility rules for inorganic salts, such as the observation that nitrates are generally soluble in , or from compiling experimental solubilities for quick lookups. For compounds, a prominent example is Yalkowsky's General Solubility (GSE), which estimates aqueous solubility based on and . The GSE is expressed as: \log S_w = 0.5 - 0.01(T_m - 25) - \log P where S_w is the molar solubility in , T_m is the in °C, and P is the . This equation has been validated on over 1,000 non-electrolytes, achieving mean absolute errors around 0.6 log units, though it assumes ideal melting behavior and performs less accurately for polar solutes. like AQUASOL and those derived from provide foundational data for such empirical models, enabling parameter fitting for specific compound classes. Theoretical methods, such as the Conductor-like Screening Model for Real Solvents (COSMO-RS), offer a physics-based approach to predict activity coefficients and thus solubility in various solvents. COSMO-RS computes chemical potentials from quantum mechanical surface charge densities obtained via calculations, accounting for electrostatic, hydrogen-bonding, and van der Waals interactions without empirical parameterization for specific systems. This model excels in multicomponent mixtures and has been applied to predict infinite dilution activity coefficients with average deviations of 0.2-0.5 log units for organic solutes in ionic liquids and aqueous systems. Its predictive power stems from treating solvents as ensembles of interacting segments, making it suitable for screening novel compounds. Computational techniques have advanced significantly with quantitative structure-activity relationship (QSAR) models and algorithms. QSAR approaches correlate solubility with molecular descriptors like molecular weight, , and topological indices, often using on datasets of drug-like compounds. These models achieve errors (RMSE) of 0.5-1.0 units on external test sets, providing interpretable insights into structural influences on solubility. In the , innovations, including graph neural networks trained on large PubChem-derived datasets exceeding 10,000 compounds, have improved accuracy for organic molecules by capturing complex three-dimensional features. Recent hybrid approaches, combining COSMO-RS with as of 2024, have further improved accuracy with RMSE values around 0.4 units for drug-like compounds. For instance, residual gated graph neural networks have reported RMSE values below 0.6 units, addressing limitations in traditional QSAR by handling diverse chemical spaces without explicit featurization. Validation of these prediction methods involves rigorous comparison against experimental data from curated databases, with performance metrics like () and RMSE quantifying accuracy across diverse compound sets. Empirical and QSAR models often underperform for outliers due to assumptions about molecular ideality, while COSMO-RS and methods show robustness but can introduce errors from quantum calculation approximations or training data biases. A key error source is polymorphism, where different forms alter and thus solubility by up to 30-fold, necessitating form-specific predictions to align with experiments. Overall, approaches combining theoretical and data-driven elements are emerging to minimize discrepancies, with ongoing benchmarks emphasizing the need for standardized validation protocols.

References

  1. [1]
    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 ...
  2. [2]
    The MSDS HyperGlossary: Solubility
    Solubility is generally expressed as the number of grams of solute in one liter of saturated solution. For example, solubility in water might be reported as 12 ...Missing: sources | Show results with:sources
  3. [3]
    Solutions
    Like Dissolves Like. By convention, we assume that one or more solutes dissolve in a solvent to form a mixture known as the solution.Missing: principle | Show results with:principle
  4. [4]
    [PDF] Solubility Rules
    Note: A substance is said to be soluble if more than 0.1 g of that substance dissolves in 100 mL solvent. Factors Affecting Solubility: I. The nature of the ...
  5. [5]
    Solubility
    Solubility is the maximum quantity of solute that can dissolve in a certain quantity of solvent or quantity of solution at a specified temperature or pressure.Missing: authoritative sources<|control11|><|separator|>
  6. [6]
    12.3 Types of Solutions and Solubility – Chemistry Fundamentals
    Temperature is another factor affecting solubility, with gas solubility typically decreasing as temperature increases (Figure 12.3.1). This inverse relation ...
  7. [7]
    [PDF] Chapter 13 notes - Web – A Colby
    The number of carbon atoms in a chain affects solubility. • The greater the number of carbons in the chain, the more the molecule behaves like a hydrocarbon. • ...
  8. [8]
    [PDF] Solubility.pdf - Colorado 4-H
    Whether a solute is soluble in a solvent depends on the chemical makeup of both the solute and the solvent, and therefore on the physical and chemical ...
  9. [9]
    William Henry: His Achievements and His Law
    The extensive and careful measurements that he made of the solubility of gases have given us the law that carries his name. Not many are familiar with his work ...<|control11|><|separator|>
  10. [10]
    Solubility Table (g/100 mL or g/100 g) Chemistry Tutorial
    A solute is usually considered to be slightly soluble, or sparingly soluble, in water if between 0.1 and 1.0 g can be dissolved in 100 mL of water. The ...
  11. [11]
    Online Catalogue Page Usage | TCI AMERICA - TCI Chemicals
    Solubility (very soluble in), The name of the solvent in which the product is soluble at a rate of > 10 g / 100 mL. Solubility (soluble in), The name of the ...
  12. [12]
    Sodium Chloride | ClNa | CID 5234 - PubChem
    6 Solubility. 36.0 g/100 g of water at 25 °C. Haynes, W.M. (ed.). CRC Handbook of Chemistry and Physics. 94th Edition. CRC Press LLC, Boca Raton: FL 2013-2014 ...
  13. [13]
    Silver Chloride | AgCl | CID 24561 - PubChem - NIH
    White solid; Darkened by light; Water solubility = 1.93 mg/L at 25 deg C; [Merck Index] Powder with lumps; [Aldrich MSDS] Chloride salts of silver are ...
  14. [14]
    The Dissolution Process – Chemistry - JMU Libraries Pressbooks
    ... intermolecular forces of attraction between solute and solvent species may prevent dissolution. ... process ... (a) ion-dipole forces; (b) dipole-dipole forces ...
  15. [15]
    Solutions - Chemistry 302
    A key idea in thinking molecularly about a solution is to imagine the intermolecular forces between the components. Since the solvent is the majority of the ...Missing: view | Show results with:view
  16. [16]
    [PDF] Solubility of ionic solids in water | Notes on General Chemistry
    Oct 18, 2007 · Solubility depends on dissolving energy (DEdissolve) and the effect of hydrated ions on water (DEorg). If total energy change is negative, the ...
  17. [17]
    Gibbs Free Energy
    The Gibbs free energy of a system at any moment in time is defined as the enthalpy of the system minus the product of the temperature times the entropy of the ...Driving Forces & Gibbs Free... · The Effect of Temperature on...
  18. [18]
    Molecular Dynamics of a Protein Surface: Ion-Residues Interactions
    Molecular dynamics simulations were carried out using as a model a few Na+ and Cl− ions enclosed in a fully hydrated simulation box with a small globular ...
  19. [19]
    CH150: Chapter 7 - Solutions - Chemistry - Western Oregon University
    Like Dissolves Like. This means that substances must have similar intermolecular forces to form solutions. When a soluble solute is introduced into a ...
  20. [20]
    [PDF] POTASSIUM NITRATE | KNO3 - PubChem - Regulations.gov
    Jun 22, 2016 · Pharmacology from NCIt. Potassium nitrate is a white to dirty gray crystalline solid. Water soluble. Noncombustible, but accelerates the.
  21. [21]
    Solubility Product Constants, K sp
    Solubility Product Constants, Ksp. Solubility product constants are used to describe saturated solutions of ionic compounds of relatively low solubility.
  22. [22]
    [PDF] Unit (5) Solutions
    Liquids that are not soluble in each other form two distinct layers (like water and oil) and are immiscible. Saturated and Unsaturated A solution whose ...
  23. [23]
    [PDF] Chapter 12. Properties of Solutions - s2.SMU
    • Mole fraction (no units); is used in calculating partial pressures of ... - mole fraction of the liquid A. X. B. - mole fraction of the liquid B. P. T. = P.
  24. [24]
    Biochemistry, Dissolution and Solubility - StatPearls - NCBI Bookshelf
    Henry's law states, “At constant temperature, the amount of gas that dissolves in a volume of liquid is proportional to the partial pressure of the gas in ...
  25. [25]
    [PDF] Henry's law constants - OSTI.GOV
    Jan 3, 2021 · To indicate the solute that a Henry's law constant refers to, it should be put in parentheses, e.g.: The Henry's law solubility constant of ...
  26. [26]
    Unleashing Carbon Dioxide - American Chemical Society
    As Henry's Law shows, the higher the pressure of a gas over a liquid, the greater the concentration of that gas in solution. To create the fizz in soda, the ...
  27. [27]
    [PDF] Lecture 3: Solubility of Gases, Liquids, and Solids in Liquids ΔG = ΔH
    to overcome the crystal lattice energy. Page 3. Of course, even if the ΔHsolution ends up slightly endothermic (like NaCl) the positive ΔSmix can make up for ...
  28. [28]
    Chapter 14 Notes - Beckford
    The ability of a solid to go into solution depends on its crystal lattice energy ... The effect of pressure is only significant for solutions of gases in liquids.
  29. [29]
    Drug Solubility: Importance and Enhancement Techniques - PMC
    Soluble, From 10 to 30. Sparingly soluble, From 30 to 100. Slightly soluble, From 100 to 1000. Very slightly soluble, From 1000 to 10,000. Practically insoluble ...
  30. [30]
  31. [31]
    [PDF] The Densities of Saturated Solutions of NaCl and KC1 from 10° to ...
    This study presents experimental data and regression coefficients for the density of saturated NaCl and KC1 solutions from 10° to 105°C.
  32. [32]
    [PDF] Major-Ion, Nutrient, and Trace-Element Concentrations in the ...
    part per million (ppm) micrograms per liter (µg/L). 1 parts per billion (ppb) ... per liter; number in parenthesis is the U.S. Environmental Protection Agency.
  33. [33]
    [PDF] issue paper on the environmental chemistry of metals
    Jan 25, 2005 · concentrations in soils exceed 9 parts per million (ppm) in a 100-kilometer-wide, 1,000- kilometer-long belt that extends southwest from New ...
  34. [34]
    [PDF] MIT Open Access Articles Predicting Solubility Limits of Organic ...
    the density used for conversion of the units should be the density of the solute and solvent mixture. However, those densities are in most cases not ...
  35. [35]
    Potassium nitrate - Sciencemadness Wiki
    Potassium nitrate ; Solubility in water. 13.3 g/100 ml (0 °C) 31.6 g/100 ml (20 °C) 246 g/100 ml (100 °C) ; Solubility, Soluble in ammonia, glycerol. Slightly ...
  36. [36]
    III. Experiments on the quantity of gases absorbed by water, at ...
    Experiments on the quantity of gases absorbed by water, at different temperatures, and under different pressures. William Henry.
  37. [37]
    Hemoglobin and Oxygen Transport Charles L
    Henry's Law: O2 dissolved (mL O2/dL) = O2 sol. (mL O2/dL per mm Hg) * PO2 (mm Hg). dissolved O2 is a linear function of PO2 · O2 solubility in blood is very low, ...
  38. [38]
    Henry's Law Constants
    Henry's law was formulated by the English chemist William Henry in the early 19th century. ... Here, the Henry's law solubility constant H s cp is used. It is ...
  39. [39]
    [PDF] On the effects of pressure on the solubility of solids in liquids
    the solubility of the solid will be raised by pressure, as in the cases of potassium sulphate or sodium chloride in water. It will be seen, therefore, that in ...
  40. [40]
    How does the high pressure affects the solubility of the drug within ...
    Moreover, our results clearly indicate that with increasing pressure the solubility of the drug within the polymer matrix is decreasing at isothermal conditions ...Missing: effects | Show results with:effects
  41. [41]
    Nitrogen Narcosis In Diving - StatPearls - NCBI Bookshelf - NIH
    Further increases in the partial pressure of nitrogen in the blood from descending deeper lend to impairments in manual dexterity and further mental decline ...
  42. [42]
    13.2 Solubility and Molecular Structure
    The ion–dipole interactions between Li + ions and acetone molecules in a solution of LiCl in acetone are shown in Figure 13.6 "Ion–Dipole Interactions in the ...Missing: view | Show results with:view
  43. [43]
    Solutions - Highland Community College
    Because the solubility of most solids increases with increasing temperature, a saturated solution that was prepared at a higher temperature usually contains ...
  44. [44]
    The Solution Process
    Generally speaking only certain molecules will dissolve in water to begin with. The old phrase "like dissolves like" or "birds of a feather flock together ...
  45. [45]
    Intermolecular Forces - EdTech Books
    How to Learn It: · Ion-Ion intermolecular forces occur when two ions interact. · Ion-Dipole intermolecular forces are between a charged ion and a polar molecule.Missing: hierarchy | Show results with:hierarchy
  46. [46]
    Intermolecular Forces – Organic Chemistry: Fundamental Principles ...
    A general rule for solubility is summarized by the expression “like dissolves like”. This means that one substance can dissolve in another with a similar ...
  47. [47]
    IMFs: Intermolecular Forces - gchem
    Define the three major intermolecular forces (IMFs) that can exist in condensed phases: dipole-dipole, hydrogen bonding, and dispersion. Predict the types of ...Missing: hierarchy | Show results with:hierarchy<|separator|>
  48. [48]
    Solvent Physical Properties
    Solvent, Polarity Index, Dipole moment, Dielectric Constant, Boiling Point, Freezing Point. Pentane, 0.0, 0, 1.84, 36.1, -129.7.Missing: solubility | Show results with:solubility
  49. [49]
    Water: Properties and Interactions - Chembook
    These intermolecular forces (IMFs) are present pretty much everywhere. ... Yes, oil will completely separate out and float as a separate layer ontop of water.
  50. [50]
    Avoid Common Pitfalls when using Henry's Law | NIST
    Sep 3, 2007 · Avoid Common Pitfalls when using Henry's Law ; Author(s). Allan H. Harvey, Francis L. Smith ; Citation. Chemical Engineering Progress ; Volume. 103.
  51. [51]
    [PDF] Avoid Common Pitfalls When Using Henry's Law
    Clearly, the use of a Henry's constant that was derived at 25°C at a different temperature could lead to serious design errors. Even a variation as small as 10 ...
  52. [52]
    Oxygen - Solubility in Fresh and Sea Water vs. Temperature
    Normally dissolved oxygen from air in fresh water and sea water (salt water) at pressures ranging 1 - 4 bar abs are indicated in the diagrams and tables below.
  53. [53]
    Probing Solubility and pH of CO2 in aqueous solutions
    As expected, the solubility of CO2 is highest in the DW and lowest in SW. The increase in salinity reduces the solubility. It should be noted that the salinity ...
  54. [54]
    4. Dissolved Oxygen - CHEMICAL FEATURES OF WATER
    At 20°C oxygen solubility content increases from a value of 9.08 mg/L at the surface to 9.98 mg/L at m depth (Table II). Proportionate reductions in values ...
  55. [55]
    Ocean acidification | National Oceanic and Atmospheric Administration
    Sep 25, 2025 · The ocean absorbs about 30% of the carbon dioxide (CO2) that is released in the atmosphere. As levels of atmospheric CO2 increase from human ...
  56. [56]
    Climate change indicators and impacts worsened in 2020
    Apr 19, 2021 · The ocean absorbs around 23% of the annual emissions of anthropogenic CO2 into the atmosphere and acts as a buffer against climate change.
  57. [57]
    Solubility Rules - Harper College
    1. Alkali metal (Group IA) compounds are soluble. · 2. Ammonium (NH4+) compounds are soluble. · 3. Nitrates (NO3-), chlorates (ClO3-), and perchlorates (ClO4-) ...
  58. [58]
  59. [59]
    Relative Strengths of Acids and Bases – Chemistry
    ... solubility in both strong acids and strong bases. In strong bases, the relatively insoluble hydrated aluminum hydroxide, Al(H2O)3(OH)3, is converted into ...Missing: curve | Show results with:curve<|separator|>
  60. [60]
    1.6: Physical properties of organic compounds - Chemistry LibreTexts
    Aug 12, 2019 · Any functional group that can donate a hydrogen bond to water (eg. alcohols, amines) will significantly contribute to water solubility. Any ...Solubility · Summary of factors... · Exercise 2.30 · Boiling point and melting point
  61. [61]
    CH105: Chapter 9 - Organic Compounds of Oxygen - Chemistry
    As the length of the chain increases, however, the solubility of alcohols in water decreases; the molecules become more like hydrocarbons and less like water.
  62. [62]
    Solubility of Organic Compounds - Chemistry Steps
    As a rough estimate, remember that water-soluble organic compounds must have an oxygen or nitrogen-containing functional group. In any case, if the molecule has ...
  63. [63]
    n-Octanol/Water Partition Coefficient (Kow/logKow)
    Jan 13, 2016 · LogKow are generally inversely related to water solubility and directly proportional to molecular weight of a substance. Kow and logKow ...
  64. [64]
    A New Approach on Estimation of Solubility and n-octanol/water ...
    The aqueous solubility (logW) and n-octanol/water partition coefficient (logPOW) are important properties for pharmacology, toxicology and medicinal chemistry.
  65. [65]
    4.4 Solubility - Chemistry LibreTexts
    Jun 5, 2019 · The longer the carbon chain in an alcohol is, the lower the solubility in polar solvents and the higher the solubility in nonpolar solvents.
  66. [66]
    Physical Properties of Alcohols and Phenols | CK-12 Foundation
    Sep 30, 2025 · Consequently, the solubility in water decreases rapidly. For instance, butan-1-ol is only moderately soluble, while hexan-1-ol is almost ...
  67. [67]
    Ethanol Effects on Apparent Solubility of Poorly Soluble Drugs in ...
    May 31, 2012 · In this work we investigated the effect of ethanol on the apparent solubility and dissolution rate of poorly soluble compounds in simulated intestinal fluid.
  68. [68]
    [PDF] Review of the cosolvency models for predicting solubility of drugs in ...
    The cosolvency, mixing a permissible non- toxic organic solvent with water, is the most common technique to increase the aqueous solubility of drugs. The common ...
  69. [69]
    Evaluation of Green and Biobased Solvent Systems for the ...
    Jan 27, 2025 · We evaluated a two-step process for β-carotene and lipid isolation from wet Rhodosporidium toruloides biomass using biphasic green and biobased solvent systems.
  70. [70]
    Green solvents in biomass delignification for fuels and chemicals
    Green solvents are advantageous over petrochemical solvents because they are less toxic and fully biodegradable. Conventional organic solvents may be replaced ...
  71. [71]
    5.4: Alloys and Crystal Defects - Chemistry LibreTexts
    Oct 14, 2023 · A solid solution can be formed by two mechanisms: (1) atom exchange or (2) interstitial mechanism. The relative size of each element in the mix ...
  72. [72]
    Solid Solutions - ASM International
    Substitutional solid solutions are those in which the solute and solvent atoms are nearly the same size, and the solute atoms simply substitute for solvent.
  73. [73]
    materials - alloys
    Alloying elements can be added to this structure in normally occupied sites to form a substitutional solid solution (b). Brass, the alloy between copper and ...
  74. [74]
    [PDF] CHAPTER 4: IMPERFECTIONS IN SOLIDS
    In interstitial solid solutions the atoms of the parent or solvent metal are bigger than the atoms of the alloying or solute metal. In this case, the ...
  75. [75]
    The freezing points, melting points, and solid solubility limits of the ...
    The freezing points, melting points, and solid solubility limits of the alloys of sliver and copper with the elements of the b sub-groups. William Hume-Rothery.
  76. [76]
    Hume-Rothery Rules and the Solid Solubility of Binary Systems
    Jun 22, 2023 · Rule 1 – Atomic Radii: A substitutional solid solution can only occur if the atomic radii of solvent and solute atoms are similar. The empirical ...Missing: primary | Show results with:primary
  77. [77]
    Phase diagrams 2 - eutectic reactions - DoITPoMS
    In the phase diagram there is a horizontal (isothermal) line at the eutectic temperature which links the equilibrium compositions of the solids. This is the ...
  78. [78]
    First-principles study of solid-solution hardening in steel alloys
    Solid solution strengthening, due to dislocation pinning by impurities, is an effective route to enhance the intrinsic hardness of alloys. In the present work, ...
  79. [79]
    What Is Solid Solution Hardening and How Does It Improve Strength?
    Jan 1, 2024 · Solid solution hardening is simply the act of dissolving one metal into another, similar to dissolving sugar into coffee.
  80. [80]
    Lattice-matched antiperovskite-perovskite system toward all-solid ...
    Aug 9, 2025 · Inorganic solid electrolytes have emerged as promising candidates for realizing all-solid-state batteries because they eliminate flammable, ...
  81. [81]
    Dissolution and pharmacokinetics of a novel micronized aspirin ...
    Micronized aspirin tablets dissolve significantly faster over a pH range from 1.2 to 6.8 compared to regular 500 mg aspirin tablets. Plasma concentration time ...
  82. [82]
    Effect of pH, Ionic Strength and Agitation Rate on Dissolution ... - NIH
    The thickness of the surface diffusion layer is inversely correlated with the dissolution rate and its depletion is expected to increase the drug diffusion ...
  83. [83]
    Correlating quartz dissolution kinetics in pure water from 25 to 625°C
    An Arrhenius expression describing a kinetic dissolution rate constant with an average activation energy of 89 ± 5 kJ/mol was regressed from the set of ...
  84. [84]
    Solvation - an overview | ScienceDirect Topics
    The solid dissolution process generally involves two sequential steps. The first step is the interaction between solid and solvent molecules to form solvated ...
  85. [85]
    4.5 Mineral Dissolution and Precipitation – Groundwater Microbiology
    Congruent dissolution refers to minerals that dissolve completely into their constituent ions, whereas incongruent dissolution applies to minerals that ...
  86. [86]
    (PDF) The dissolution and conversion of Gypsum and Anhydrite
    Aug 6, 2025 · This chapter is an attempt to summarise the present knowledge of the dissolution chemistry and kinetics of gypsum and anhydrite.
  87. [87]
    Steady-state dissolution kinetics of aluminum-goethite in ... - PubMed
    ... Al did not scale with those of Fe, indicating incongruent dissolution. ... These results are consistent with the presence of Al(OH)3 inclusions in Al-goethite.
  88. [88]
    Weathering Processes and Mechanisms Caused by Capillary ...
    This geochemical study suggests salts precipitate via incongruent reactions rather than congruent precipitation, where hexahydrite is the precursor and limiting ...
  89. [89]
    A thermodynamic model of dissolution and precipitation of calcium ...
    Several models have been proposed for the incongruent dissolution of calcium silicate hydrate (C-S-H) gel, which is the principal product of hydrated cement ...
  90. [90]
    Incongruent dissolution of silicates and its impact on the environment
    Dec 18, 2023 · The paper analyzes the process of incongruent dissolution of silicates taking place in close proximity to a talc mine.Missing: hydration | Show results with:hydration
  91. [91]
    50 Years with solubility parameters—past and future - ScienceDirect
    The solubility parameter was first introduced by Hildebrand and Scott in 1950, just over 50 years ago [1], [2]. The Hildebrand solubility parameter, δt, is ...Abstract · Solvent-Reducible Coatings · Pigment Surface Optimization
  92. [92]
    18.1: Solubility Product Constant, Ksp - Chemistry LibreTexts
    Feb 13, 2024 · The equilibrium constant for a dissolution reaction, called the solubility product (Ksp), is a measure of the solubility of a compound.The Solubility Product · Note · Example 18 . 1 . 1 · Example 18 . 1 . 2
  93. [93]
    Ksp Table
    Silver chloride AgCl 1.8×10–10. Silver chromate Ag2CrO4 1.1×10–12. Silver cyanide AgCN 1.2×10–16. Silver iodate AgIO3 3.0×10–8. Silver iodide AgI 8.5×10–17.
  94. [94]
    Common Ion Effect - Chemistry LibreTexts
    May 4, 2025 · Adding a common ion decreases solubility, as the reaction shifts toward the left to relieve the stress of the excess product. Adding a common ...Introduction · Simple Example · Common Ion Effect with Weak...
  95. [95]
    18.5: Criteria for Precipitation and its Completeness
    Jul 12, 2023 · If the concentrations are such that Q is less than Ksp, then the solution is not saturated and no precipitate will form. We can compare ...<|separator|>
  96. [96]
    The change in solubility of calcium carbonate with temperature and ...
    The temperature derivative of solubility of calcite at constant CO2 concentration varies from −10−6 to −3 × 10−5 molal/°C, being a maximum for the highest CO2 ...
  97. [97]
    [PDF] "Historic Papers in Electrochemistry"
    Dehye u. Huckel, Zur Theorie der Elektrolyte. Physik. Zeitschr. XXIV, 1923. Ionenwirkungen. Dementsprechend zerlegen wir. U in zwei Bestandteile ...
  98. [98]
    SOLUBILITY. XII. REGULAR SOLUTIONS 1 - ACS Publications
    This article, published in 1929, is about solubility and regular solutions. It has 385 citations.
  99. [99]
    Group‐contribution estimation of activity coefficients in nonideal ...
    A group-contribution method is presented for the prediction of activity coefficients in nonelectrolyte liquid mixtures. The method combines the solution-of ...
  100. [100]
    Analysis of the UNIFAC-Type Group-Contribution Models at the ...
    An analysis of the behavior of UNIFAC-type group-contribution models at the infinite-dilution region is presented. It is shown that the modified UNIFAC and ...
  101. [101]
    Advances in Machine Learning Approaches for Predicting Aqueous ...
    Mar 21, 2025 · This study aims at providing a detailed review on the prediction of aqueous solubility in drug discovery, using machine learning approaches and ...
  102. [102]
    Prediction of parameters of group contribution models of mixtures by ...
    This new approach is applied to UNIFAC, an established group contribution method for predicting activity coefficients in mixtures. The developed MCM yields ...Missing: original | Show results with:original
  103. [103]
    Solubility advantage of amorphous pharmaceuticals - PubMed
    Conclusion: It has been demonstrated that the theoretical approach does provide an accurate estimate of the maximum solubility enhancement by an amorphous drug ...
  104. [104]
    [PDF] ENHANCEMENT OF CARBAMZEPINE SOLUBILITY USING ...
    In the literature, solid dispersions have shown tremendous potential for improving drug solubility and dissolution. In this study the main objective was to ...
  105. [105]
    Comprehensive analysis of heavy metal soil contamination in ...
    Acidic soils typically have pH levels below 7, which can increase the solubility of heavy metal ions in the soil, consequently leading to higher levels of ...
  106. [106]
    Bio-chelate assisted leaching for enhanced heavy metal ... - Nature
    Jun 20, 2024 · The pH level of the solution is believed to impact the solubility of heavy metals, as well as the structure and charge of chelators, and their ...
  107. [107]
    Clean-Up of Heavy Metals from Contaminated Soil by ...
    May 2, 2023 · The removal of heavy metal defilements can be accomplished using phytoremediation techniques, including phytoextraction, phytovolatilization, ...
  108. [108]
    Waste Water Treatment - Chemistry LibreTexts
    Mar 11, 2023 · The main goal of the precipitation portion of the wastewater treatment process is to remove soluble metal ions and phosphates from water.Precipitation Reactions in... · Removal of Fe2+ and Mn2+ · Removal of Phosphates
  109. [109]
    [PDF] On the design of crystallization-based separation processes - CEPAC
    Jan 11, 2006 · An analysis is given of various aspects of the design of crystallization-based separation processes. A literature review is included for ...
  110. [110]
    How to separate KNO3 and NaCl solution - Quora
    Dec 18, 2024 · If it is a solution in water you can separate the potassium chloride by fractional crystallization, since the solubility of the potassium salt ...How can we separate a mixture of KNO3 and KCl by using fractional ...After mixing KNO3 and KCl, both saturated in solution, will the filter ...More results from www.quora.com
  111. [111]
    Hansen Solubility Parameters Applied to the Extraction of ... - NIH
    Aug 21, 2023 · The selection of extraction solvents is generally based on the solubility difference between target analytes and the undesired matrix components ...
  112. [112]
    Solvent Extraction Techniques - Organomation
    In a typical extraction, two immiscible liquids are mixed, and the compound of interest partitions between the two phases based on its relative solubility. The ...
  113. [113]
    Unifying Future Ocean Oxygen Projections Using an Oxygen Water ...
    Apr 28, 2025 · Climate change reduces ocean oxygen levels, posing a serious threat to marine ecosystems and their benefits to society.
  114. [114]
    Prediction of Aqueous Solubility of Organic Compounds by the ...
    This manuscript compares the general solubility equation (GSE) of Jain and Yalkowsky5 with four different methods to test 21 compounds described by ...
  115. [115]
    Prediction of Drug Solubility by the General Solubility Equation (GSE)
    Aug 5, 2025 · The revised general solubility equation (GSE) proposed by Jain and Yalkowsky is used to estimate the aqueous solubility of a set of organic nonelectrolytes.Missing: paper | Show results with:paper
  116. [116]
    An overview of the performance of the COSMO-RS ... - RSC Publishing
    This paper evaluates COSMO-RS for predicting infinite dilution activity coefficients of molecular solutes in ionic liquids, using 41,868 data points.
  117. [117]
    QSAR-based solubility model for drug-like compounds - ScienceDirect
    This paper describes a QSAR solubility model based on freely available information and robust enough to correctly identify soluble drug-like compounds in ...
  118. [118]
    SolPredictor: Predicting Solubility with Residual Gated Graph Neural ...
    Jan 5, 2024 · The proposed SolPredictor is designed with the aim of developing a computational model for solubility prediction.
  119. [119]
    Prediction of the water solubility by a graph convolutional-based ...
    Apr 21, 2025 · We developed a graph convolutional neural network model (GNN) to predict the water solubility in the form of log S w based on a highly curated dataset of 9800 ...Solubility Dataset · Neural Networks · Curation Of The Dataset
  120. [120]
    Accurate Measurement, and Validation of Solubility Data - PMC - NIH
    Finally, this study provides guidelines to accurately measure and validate solubility data for polymorphic compounds, particularly when they are prone to ...
  121. [121]
    Physics-Based Solubility Prediction for Organic Molecules
    Solubility represents the point at which a stable solute–solvent thermodynamic equilibrium has been achieved, with the rate of dissolution equal to the rate of ...