Fact-checked by Grok 2 weeks ago

Partition coefficient

The partition coefficient, often denoted as P, is a key physicochemical parameter that describes the equilibrium distribution of a neutral solute between two immiscible phases, such as an organic solvent and water, defined as the ratio of the solute's concentration in the organic phase (C_\text{org}) to that in the aqueous phase (C_\text{aq}) at equilibrium: P = C_\text{org} / C_\text{aq}. This unitless value is frequently expressed in logarithmic form as \log P, with the n-octanol/ partition coefficient (\log K_\text{ow}) serving as the standard metric for assessing a compound's , or affinity for versus . In and pharmaceutical sciences, \log P is indispensable for evaluating a molecule's potential pharmacokinetic behavior, including across biological membranes, , and to plasma proteins, where optimal values typically range from -1 to 5 to satisfy and enhance oral . High \log P values indicate greater hydrophobicity, which can improve membrane permeability but may reduce aqueous and increase toxicity risks through . Beyond pharmacology, the partition coefficient informs environmental fate modeling by predicting how chemicals partition between media like air, water, soil, and biota; for instance, \log K_\text{ow} correlates with a pollutant's tendency to sorb onto organic matter in sediments or accumulate in fatty tissues, influencing remediation strategies and regulatory assessments. Related variants, such as the distribution coefficient (D) or soil organic carbon-water partition coefficient (K_\text{oc}), account for ionization or matrix effects, extending its utility in toxicology and ecotoxicology.

Fundamentals

Definition and Nomenclature

The concept of the partition coefficient emerged from early 20th-century investigations into solute distribution between immiscible solvents, with foundational work by German physical chemist Walther Nernst in 1891. Nernst's distribution law described how a solute partitions itself such that the ratio of its concentrations in the two phases remains constant at equilibrium, laying the groundwork for quantitative analysis of phase separations in physical chemistry. The partition coefficient, denoted as P, is precisely defined as the ratio of the equilibrium concentrations of a neutral solute species in two immiscible phases: P = \frac{[\text{solute}]_{\text{phase 1}}}{[\text{solute}]_{\text{phase 2}}} where the concentrations are typically expressed in units. This measure applies specifically to un-ionized, molecular , as ionized forms generally exhibit lower partitioning into non-polar phases due to electrostatic interactions with the aqueous . In ideal dilute solutions, the definition assumes equal activity coefficients across phases, permitting the direct use of concentration ratios to approximate thermodynamic activities without correction for non-ideality. Nomenclature for this property includes variations such as and , though the International Union of Pure and Applied Chemistry (IUPAC) discourages "" as a for these terms to avoid ambiguity. The (D), in contrast, represents the of total analytical solute concentrations (encompassing all ) and varies with conditions like or complexation, distinguishing it from the invariant partition coefficient for a specific . coefficient serves as a for in contexts. The partition coefficient is inherently dimensionless, reflecting its nature as a ; for practical handling of values spanning orders of magnitude, it is commonly logarithmically transformed to log [P](/page/P′′).

Partition Coefficient and Log P

The partition coefficient P specifically applies to the partitioning of undissociated, molecules between two immiscible phases, most commonly n-octanol and , at . It is defined by the equation P = \frac{C_{\text{octanol}}}{C_{\text{water}}} where C_{\text{octanol}} and C_{\text{water}} represent the equilibrium concentrations of the neutral solute in the octanol and phases, respectively. This ratio quantifies the relative affinity of the compound for the lipophilic octanol phase versus the aqueous phase. To facilitate comparisons and modeling across compounds with vastly differing solubilities, the partition coefficient is typically expressed on a logarithmic scale as \log P = \log_{10} P. This transformation compresses the wide dynamic range of P values (often spanning several orders of magnitude) into a more manageable numerical scale, enabling its use as a key physicochemical descriptor. The \log P value thus serves as a primary measure of , with positive values (\log P > 0) indicating a preference for the octanol phase and greater lipophilicity, while negative values (\log P < 0) signify higher hydrophilicity and affinity for water; a value of \log P = 0 denotes equal partitioning between the phases. The adoption of \log P gained prominence in the 1960s through quantitative structure-activity relationship (QSAR) studies, particularly the seminal work of Hansch and Fujita, who introduced it as a linear free-energy parameter to correlate molecular hydrophobicity with biological potency. Their \rho-\sigma-\pi analysis framework demonstrated how \log P could predict drug transport and activity, establishing it as a foundational metric in medicinal chemistry and environmental toxicology. However, \log P is strictly applicable to neutral, non-ionizable compounds under conditions where dissociation does not occur, as it assumes a single molecular species without pH-dependent effects.

Distribution Coefficient and Log D

The distribution coefficient, denoted as D, represents the ratio of the total concentrations of an analyte (encompassing both its neutral and ionized forms) in two immiscible phases, typically n-octanol and water, at a specified pH. Unlike the partition coefficient P, which applies solely to the neutral species, D accounts for pH-dependent speciation, making it essential for ionizable compounds. The logarithmic form, \log D = \log_{10} D, quantifies this distribution and varies with pH. For ionizable compounds, \log D derives from \log P adjusted for the fraction of neutral species, assuming negligible partitioning of ionized forms into the organic phase. Consider a monoprotic acid HA in equilibrium with its conjugate base A^-: \text{HA} \rightleftharpoons \text{A}^- + \text{H}^+, \quad K_a = \frac{[\text{A}^-][\text{H}^+]}{[\text{HA}]}, \quad \text{p}K_a = -\log_{10} K_a. In the aqueous phase, the total concentration is [\text{HA}]_\text{aq} + [\text{A}^-]_\text{aq} = [\text{HA}]_\text{aq} (1 + 10^{\text{pH} - \text{p}K_a}), since [\text{A}^-]_\text{aq}/[\text{HA}]_\text{aq} = 10^{\text{pH} - \text{p}K_a}. In the octanol phase, only the neutral HA partitions significantly, so the total is [\text{HA}]_\text{oct}. Thus, D = \frac{[\text{HA}]_\text{oct}}{[\text{HA}]_\text{aq} (1 + 10^{\text{pH} - \text{p}K_a})} = \frac{P}{1 + 10^{\text{pH} - \text{p}K_a}}, where P = [\text{HA}]_\text{oct}/[\text{HA}]_\text{aq}. Taking the base-10 logarithm yields \log D = \log P - \log(1 + 10^{\text{pH} - \text{p}K_a}). For a monoprotic base B in equilibrium with its conjugate acid BH^+ (where pKa refers to BH^+): \text{BH}^+ \rightleftharpoons \text{B} + \text{H}^+, \quad K_a = \frac{[\text{B}][\text{H}^+]}{[\text{BH}^+]}, \quad \text{p}K_a = -\log_{10} K_a. The total aqueous concentration is [\text{B}]_\text{aq} + [\text{BH}^+]_\text{aq} = [\text{B}]_\text{aq} (1 + 10^{\text{p}K_a - \text{pH}}), since [\text{BH}^+]_\text{aq}/[\text{B}]_\text{aq} = 10^{\text{p}K_a - \text{pH}}. In octanol, only neutral B partitions, so D = \frac{[\text{B}]_\text{oct}}{[\text{B}]_\text{aq} (1 + 10^{\text{p}K_a - \text{pH}})} = \frac{P}{1 + 10^{\text{p}K_a - \text{pH}}}, and \log D = \log P - \log(1 + 10^{\text{p}K_a - \text{pH}}). These relations stem from the governing speciation. The pKa value modulates partitioning by determining the pH range over which ionization occurs, with \log D approaching \log P when the species is predominantly neutral (low pH for acids, high pH for bases) and decreasing sharply near the pKa. Graphical representations of \log D versus pH exhibit sigmoidal curves: for acids, \log D remains constant at low pH, then declines linearly with a slope of approximately -1 over a 2-unit pH interval centered on pKa, before plateauing at higher pH; bases show the inverse, with a rise at higher pH. This pH dependence highlights how environmental or physiological conditions alter effective lipophilicity. In real-world scenarios, such as biological fluids (e.g., blood at pH 7.4 or gastrointestinal tract varying from pH 1.5 to 8), pH influences molecular speciation, thereby affecting partitioning and bioavailability of ionizable drugs. For instance, at physiological pH, many drugs exist partly ionized, reducing their membrane permeability compared to neutral forms. Consequently, \log D provides a more accurate assessment of lipophilicity than \log P for ionizable pharmaceuticals, enabling better predictions of absorption, distribution, and toxicity under relevant conditions.

Theoretical Foundations

Equilibrium Partitioning

Partitioning of a solute between two immiscible phases is a reversible equilibrium process first formalized by the in 1891, which posits that the ratio of solute concentrations in the two phases remains constant at a given temperature when the solute is distributed without chemical reaction or association. This law laid the foundation for understanding partitioning as an equilibrium phenomenon driven by the solute's affinity for each phase. Thermodynamically, the partition coefficient P relates to the standard Gibbs free energy change \Delta G^\circ for the transfer of the solute from one phase to the other via the equation \Delta G^\circ = -RT \ln P, where R is the gas constant and T is the absolute temperature (typically 298 K or 25°C); a negative \Delta G^\circ indicates spontaneous partitioning favoring the second phase. This equilibrium arises from the balance of enthalpic and entropic contributions, with \Delta G^\circ = \Delta H^\circ - T \Delta S^\circ, where \Delta H^\circ reflects solvation energies and \Delta S^\circ accounts for changes in solute-solvent interactions. Key factors influencing partitioning include entropy gains or losses from solute-solvent ordering, such as hydrophobic effects in aqueous systems, and enthalpic terms from specific interactions like hydrogen bonding or cavity formation in the nonpolar phase. For instance, polar solutes with strong hydrogen bonding to water exhibit lower partition coefficients into nonpolar solvents due to unfavorable enthalpic penalties upon desolvation. For ionizable compounds, pH-dependent partitioning is addressed via the distribution coefficient (see Fundamentals section). In ideal systems, partitioning follows with P independent of concentration, but real solutions often deviate due to non-ideal behavior captured by \gamma. The thermodynamic partition coefficient is P = \frac{c_\text{org} \gamma_\text{org}}{c_\text{aq} \gamma_\text{aq}} (ratio of activities), while the apparent (concentration-based) partition coefficient is K = \frac{c_\text{org}}{c_\text{aq}} = P \times \frac{\gamma_\text{aq}}{\gamma_\text{org}}, where \gamma \neq 1 arises from solute-solute or solute-solvent interactions. Salting-out effects, where added electrolytes reduce solute solubility in the aqueous phase by altering water structure and increasing , enhance partitioning into the organic phase, as quantified by the \log (S_0 / S) = k_s C_s with salting constant k_s > 0. Extensions to multi-phase systems, such as three-phase liquid-liquid extractions or emulsions, involve sequential partitioning equilibria, where the overall distribution depends on pairwise coefficients between phases, including interfacial effects in emulsions that can trap solutes and alter effective P. In emulsions, influence partitioning by modifying interfacial tension and solute adsorption, leading to non-equilibrium distributions in dynamic systems.

Relation to Physicochemical Properties

The partition coefficient, particularly its logarithmic form log P, exhibits strong empirical correlations with aqueous (S_w), reflecting the balance between hydrophobic and hydrophilic interactions in a . A common approximation for non-polar compounds is log P ≈ -log S_w + constant, where S_w is the aqueous in units; for hydrocarbons, this is more precisely expressed as log P = 5.00 - 0.67 log S_w. This relation arises because higher (larger log P) typically reduces in by favoring partitioning into non-aqueous phases. However, limitations include deviations for polar or ionizable compounds, where the slope may shift due to specific solute-solvent interactions, and the constant varies with molecular class, reducing accuracy beyond simple hydrocarbons. The Abraham general solvation model provides a structured framework linking log P to key physicochemical properties, expressed as log P = c + eE + sS + aA + bB + vV, where E represents excess molar refraction (related to ), S denotes dipolarity/, A and B are hydrogen-bond acidity and basicity (quantifying H-bond donors and acceptors), and V is the McGowan characteristic volume (a proxy for ). (V) contributes to the , increasing log P as molecular size grows, while —often correlated with A and B—inversely affects log P by enhancing interactions and reducing . This model, validated across diverse solvents, highlights how increases in V or decreases in polar features elevate log P, aiding in property predictions for complex molecules. Log P serves as a predictor for several related properties, influencing environmental and material behaviors. For factors (BCF) in organisms, log BCF often correlates linearly with log P up to values around 6, with BCF ≈ 0.5 × 10^{log P} for many organics, as higher promotes in fatty tissues. decreases with increasing log P in , as lipophilic compounds exhibit stronger intermolecular forces, reducing ; this is captured in regressions like log VP = f(-log P, other terms) for environmental fate modeling. Similarly, log P correlates positively with in non-polar series, where larger, more hydrophobic molecules form tighter crystals, though this weakens for polar substituents. In like n-alkanes, log P demonstrates a linear with chain length, with an incremental increase of approximately 0.5 log units per , illustrating the additive nature of alkyl chains in enhancing partitioning. Recent advancements since 2020 have leveraged to refine these correlations, deriving nonlinear links between log P, , and properties like from large datasets. For instance, graph neural networks such as Chemprop have improved predictions of log P by incorporating 2D/ molecular descriptors, achieving mean absolute errors around 0.44 log units for diverse compounds and revealing hidden interactions overlooked in linear models.

Measurement Methods

Shake-Flask and Separating-Funnel Techniques

The shake-flask technique represents a foundational experimental approach for directly measuring the n-octanol/ partition coefficient (P_{ow}), defined as the of the concentrations of a solute in the octanol and aqueous phases (P_{ow} = [solute]{octanol} / [solute]{}). To perform the procedure, analytical-grade n-octanol and high-purity are first mutually saturated by vigorous shaking in equal volumes for at least 24 hours, followed by gravity or centrifugal separation to obtain the pre-equilibrated phases. A known mass or concentration of the test substance (typically up to 0.01 mol/L total) is then introduced into a stoppered vessel, such as a glass centrifuge tube or , containing the two phases in a volume adjusted to the anticipated P_{ow}—often 1:1 (e.g., 10 mL each) for balanced distribution, but with more for hydrophilic compounds (P_{ow} < 1) or more octanol for lipophilic ones (P_{ow} > 10) to ensure detectable levels in both phases. The vessel is shaken mechanically or by hand (e.g., 100-500 oscillations over 5-10 minutes) at a controlled of 20-25°C until is reached, which generally requires 30 minutes for most compounds, though it may extend to 3 hours for slower equilibrating substances. Post-equilibration, the phases are separated by (typically 3000 rpm for 5-10 minutes) to produce a clean , with care taken to sample the aqueous phase via to avoid octanol droplets. Concentrations are quantified separately using UV-Vis for UV-active solutes, (GC) for volatiles, or (HPLC) for general applicability, yielding P_{ow} from the of measured concentrations after correcting for any impurities or blanks. Multiple runs (at least three, with varied volume ratios like 1:1, 1:2, and 2:1) are conducted in duplicate to confirm , with acceptable log P_{ow} variation ≤ 0.3 units. The separating-funnel technique serves as a gravity-based variant of the shake-flask method, ideal for larger-scale measurements (e.g., 50-100 mL total volume) where centrifugation equipment is unavailable or sample quantities are ample. In this approach, the pre-equilibrated octanol and water phases, along with the test substance, are combined in a glass separating funnel, which is securely stoppered and inverted/shaken gently but thoroughly (e.g., 1-2 minutes initially, followed by periodic mixing) to promote distribution without excessive foaming. Equilibrium is similarly attained after about 30 minutes of intermittent agitation, after which the funnel is clamped upright to allow density-driven phase separation over 10-30 minutes, forming distinct layers (octanol on top). The lower aqueous layer is drained first via the stopcock into a collection vessel, followed by the upper octanol layer, with the interface discarded to prevent cross-contamination. Analysis proceeds as in the shake-flask method, using titration for acidic/basic solutes (e.g., acetic acid via NaOH standardization) or instrumental techniques like UV-Vis or GC for others, calculating P_{ow} from the concentration ratio adjusted for phase volumes. A key error source in this method is emulsion formation, particularly with amphiphilic or surface-active compounds, which can trap solute at the interface and bias results; mitigation involves slower shaking, temperature control, or electrolyte addition to water, though the latter may alter partitioning. Both techniques excel in their simplicity, relying on basic apparatus and providing accurate, direct P_{ow} values for non-ionizable, non-surface-active compounds within log P_{ow} -2 to +4 (extendable to +5 with care), without needing surrogate calibrants or complex modeling. For instance, they reliably quantify for pharmaceuticals like (log P_{ow} ≈ -0.07) or (log P_{ow} ≈ 2.13), establishing key physicochemical context for predictions. However, drawbacks include high labor demands for phase handling and analysis, time inefficiency for high-throughput needs (each run taking 1-2 hours plus analysis), and limitations for volatile solutes prone to evaporative loss during shaking or separation, as well as low-solubility compounds where one-phase concentrations fall below detection limits (e.g., <10^{-6} M). Emulsions and incomplete separation further compromise precision in the separating-funnel variant, potentially requiring 10-20% more replicates for statistical confidence. Standardization of these methods is outlined in OECD Test Guideline 107 (adopted 1995), which mandates analytical-grade reagents, temperature control (±0.5°C), blank corrections, and validation via mass balance (recovery 90-110%) to ensure inter-laboratory reproducibility, with the guideline drawing from earlier protocols to minimize artifacts like octanol carryover. Historically, these manual equilibration techniques originated in pre-1950s pharmacology research, where pioneers like Hans Horst Meyer and Charles Ernest Overton (circa 1899-1901) employed analogous phase distribution assays between water and oils (e.g., olive oil) to correlate solute lipophilicity with narcotic potency in tadpoles and bacteria, laying empirical groundwork for quantitative log P assessments in drug design.

Chromatographic and pH-Metric Methods

Chromatographic methods, particularly high-performance liquid chromatography (), provide an indirect approach to determining the partition coefficient () by correlating solute retention times with known partitioning behavior. In reversed-phase , the capacity factor (k), defined as k = (t_R - t_0)/t_0 where t_R is the retention time and t_0 is the dead time, is measured under standardized conditions, often using a C18 column with a methanol-water mobile phase gradient. Retention is linked to through linear solvation energy relationships (), which model intermolecular interactions via the equation log k = c + eE + sS + aA + bB + vV, where descriptors E (excess molar refraction), S (dipolarity/polarizability), A (hydrogen bond acidity), B (hydrogen bond basicity), and V () quantify solute-solvent forces, with system-specific coefficients c, e, s, a, b, and v obtained from calibration with standards of known . For ionizable compounds, modified includes ionization terms d⁺D⁺ and d⁻D⁻ to account for pH-dependent charge effects. A simplified calibration often employs log k = S log P + c, where S is the slope reflecting phase selectivity and c the intercept, derived from plotting log k against log P for reference compounds, enabling prediction of unknown log P values from measured retention. These methods are calibrated using sets of n-octanol-water log P standards, ensuring applicability across diverse chemical classes. pH-metric methods utilize potentiometric titration to assess log P by monitoring pH-dependent distribution profiles in a biphasic system, typically n-octanol-water. The technique involves titrating the compound across a pH range that spans its pKa, allowing calculation of the apparent partition coefficient (log D) at each pH via integration of the titration curve, which incorporates the to separate neutral (log P) and ionized contributions: for acids, log D = log P - log(1 + 10^{pH - pKa}), and for bases, log D = log P - log(1 + 10^{pKa - pH}). This yields log P for the neutral species by extrapolating to conditions where ionization is negligible. Both approaches offer high throughput, suitable for log P values from -2 to 6, with automation enabling hundreds of samples per day and minimal sample requirements (micrograms). They are insensitive to impurities and provide reproducible results (standard deviation <0.1 log units) when calibrated properly. Limitations include the need for representative calibration sets to assume linear free energy relationships, potential inaccuracies for highly polar or ionic compounds outside the model domain, and sensitivity to mobile phase composition in HPLC. Validation against the reference shake-flask method shows strong correlations, with R² > 0.95 in modern setups for compounds spanning log P 0 to 5. Since , integration of ultra-high-performance liquid chromatography (UHPLC) has enhanced efficiency, reducing analysis time to under 2 minutes per sample while maintaining accuracy through faster gradients and smaller particles, further expanding throughput for applications.

Electrochemical and Emerging Techniques

Electrochemical methods for measuring partition coefficients leverage ion-transfer at the interface between two immiscible electrolyte solutions (ITIES), enabling the detection of partitioning behavior for ionic or ionizable compounds without requiring . In this approach, is applied to monitor the transfer of species across the interface, typically using as the organic phase due to its electrochemical stability and similarity to n-octanol in properties. The partition coefficient is derived from the voltammetric response, such as peak potentials or half-wave potentials, which reflect the equilibrium distribution; for instance, the formal transfer potential is related to log P via the Nernstian relation for ion transfer, Δφ° = (RT/F) log P + constant terms for . Microfluidic devices offer a single-phase or low-volume alternative for rapid partition coefficient assessment, integrating liquid-liquid with on-chip detection to minimize sample use and enable high-throughput . These systems facilitate the formation of biphasic flows, such as octanol-water, allowing partitioning in microliter volumes; impedance can then probe the changes associated with solute between phases, providing monitoring of concentration ratios. For example, continuous-flow microfluidic setups have demonstrated accurate log P values for diverse compounds, with times under 5 minutes per sample. Emerging techniques extend these capabilities to complex systems, including (MST), which assesses partitioning indirectly through thermophoretic mobility influenced by hydrophilicity in mixed solvents, and NMR-based methods that directly quantify solute distributions via analysis. In MST, the Soret coefficient reflects phase preferences in temperature gradients, correlating with octanol-water log P for assessments. NMR partitioning, particularly using 2D spectra, enables the determination of individual partition coefficients in multicomponent mixtures without isolation, as signal intensities yield equilibrium concentrations in each phase; this is especially valuable for biological or environmental samples where traditional methods fail due to interference. These techniques provide key advantages, including for reduced consumption (often <1 μL) and real-time data acquisition, facilitating in situ monitoring of dynamic partitioning processes. However, challenges persist, such as heightened sensitivity to impurities that adsorb at interfaces and alter voltammetric signals, potentially skewing partition estimates by up to 0.5 log units in impure samples. Post-2020 developments have integrated artificial intelligence to enhance electrochemical sensors for lipophilicity-related measurements, using machine learning to deconvolute complex voltammograms and predict partition coefficients from raw impedance or current data with improved accuracy over traditional fitting. For instance, AI algorithms process multispectral electrochemical responses to classify and quantify partitioning in noisy environments, achieving detection limits below 10^{-6} M for ionizable drugs.

Prediction Methods

Empirical and Fragment-Based Approaches

Empirical approaches to predicting the partition coefficient, often denoted as log P, rely on the additivity principle, which assumes that the lipophilicity of a molecule can be estimated by summing contributions from its constituent fragments or substituents. These methods emerged in the mid-20th century as practical tools for estimating log P values without direct experimentation, drawing from extensive compilations of measured data. Fragment-based methods, in particular, break down molecules into atomic or structural units, assigning predefined hydrophobicity values to each, while empirical substituent constants adjust base values for modifications. This approach has been foundational in , enabling rapid screening of compound libraries. One of the seminal fragment-based methods is the CLOGP algorithm, developed by Corwin Hansch and Albert Leo in the 1970s, which calculates log P as the sum of fragment constants (f_i) for each molecular piece, corrected for intramolecular interactions. For example, in ethanol (CH3CH2OH), the calculation proceeds as follows: the methyl group (CH3) contributes +0.50, the methylene group (CH2) contributes +0.50, and the hydroxyl group (OH) contributes -1.16 (accounting for hydrogen bonding effects); the total sums to approximately -0.16, close to ethanol's experimental log P of -0.31, with corrections for proximity effects improving accuracy. This method uses a dictionary of over 1,000 fragment values derived from regression against octanol-water partition data. Parallel to fragment additivity, empirical methods employ substituent constants to quantify deviations from a parent hydrocarbon's log P. Rekker's fragmental constants (f-values), introduced in the 1970s, assign values to groups like -CH3 (+0.50) or -OH (-1.16), enabling predictions via summation with a correction factor (I_m) for molecular size, as in log P = Σf_i + I_m. Similarly, the Leo-Hansch π constants measure the lipophilic increment of substituents relative to hydrogen (e.g., π for -CH3 is +0.50, for -OH is -1.16), applied in equations like log P = log P_0 + Σπ for substituted benzenes. These constants, tabulated from experimental log P measurements, allow modular predictions for analog series in drug design. Such methods are trained and validated against large databases of experimentally determined log P values, such as the , which contains approximately 13,000 compounds with measured octanol-water partition coefficients spanning diverse chemical classes. By regressing fragment contributions against this data, models achieve root-mean-square errors (RMSE) typically between 0.2 and 0.5 log units for rigid molecules, though accuracy diminishes for flexible structures with conformational variability or tautomerism, where intramolecular hydrogen bonding or entropy effects are not fully captured. Limitations include overestimation for highly polar or ionic species and the need for manual corrections in complex cases. Historically, these approaches evolved from manual tables in the 1970s, such as Hansch and Leo's pioneering work, to integrated software tools like (now part of ACD/Percepta), which automate fragment assignment and apply electronic corrections using over 7,000 fragments. This progression has made empirical predictions accessible for high-throughput applications while maintaining reliance on curated experimental datasets for reliability.

Computational and Knowledge-Based Models

Computational and knowledge-based models for predicting partition coefficients, particularly the octanol-water partition coefficient (), leverage advanced algorithms and molecular representations to achieve high accuracy without relying on experimental measurements. These approaches integrate machine learning, quantitative structure-activity relationship () techniques, and quantum mechanical calculations to model the underlying physicochemical interactions driving partitioning behavior. By training on extensive datasets of experimental values, these models enable rapid screening of chemical libraries in drug discovery and environmental assessments. Knowledge-based models, such as neural networks trained on large experimental datasets, have become prominent for logP prediction due to their ability to capture nonlinear relationships in molecular data. For instance, tools like ChemAxon's CXLogP employ artificial neural networks (ANNs) that achieve root-mean-square errors (RMSE) as low as 0.31 on blind test sets from the SAMPL6 challenge, outperforming traditional methods on diverse organic compounds. Similarly, Schrödinger's software suite incorporates machine learning-enhanced predictions, often integrating neural networks with molecular descriptors for robust performance across ionizable species. These models typically use datasets exceeding 10,000 compounds, enabling generalization to novel structures while maintaining low prediction errors under 0.5 log units. Atom-based and 3D-QSAR methods extend these predictions by incorporating spatial and topological molecular features, such as the topological polar surface area (TPSA), which correlates with hydrogen bonding potential and solvation effects. Comparative Molecular Field Analysis (CoMFA), a foundational 3D-QSAR technique, generates steric and electrostatic field descriptors around aligned molecular conformations to build partial least squares (PLS) regression models for logP. In CoMFA applications, TPSA serves as a key descriptor, with models demonstrating cross-validated correlation coefficients (q²) above 0.7 when trained on congeneric series of pharmaceuticals. These approaches are particularly valuable for understanding how molecular shape influences partitioning, as validated in reviews of QSAR tools for physicochemical properties. Quantum mechanical methods provide a physics-based alternative by computing solvation free energies (ΔG_solv) to derive logP values, often through implicit solvation models. The Solvation Model based on Density (SMD) calculates ΔG_solv by accounting for short-range and long-range solute-solvent interactions using quantum chemical charge densities, allowing logP estimation via the relation logP ≈ -(ΔG_solv,water - ΔG_solv,octanol)/(2.303 RT). SMD, parameterized for over 100 solvents, yields RMSE values around 0.6-0.8 log units for neutral organics when combined with density functionals like . Complementarily, the Conductor-like Screening Model (COSMO) derives surface charge densities from quantum calculations to predict activity coefficients and thus partitioning, with applications showing improved accuracy for polar compounds over classical force fields. Recent advances in the 2020s have introduced graph neural networks (GNNs) to represent molecules as graphs, where atoms are nodes and bonds are edges, enabling end-to-end learning of logP from structural data. GNN models, such as those using message-passing architectures, have achieved RMSE below 0.4 on large datasets like PubChem, surpassing earlier QSAR by directly learning featurizations. For example, multi-fidelity GNNs integrate low- and high-accuracy quantum data to predict partition coefficients with enhanced precision for underrepresented chemical spaces. These developments address gaps in traditional ML by handling molecular symmetry and stereochemistry natively. As of 2025, further progress includes participation in the SAMPL9 LogP challenge, which tested predictions on 16 drug-like molecules using distributed computing like , achieving improved RMSE for complex structures, and the integration of large language models (LLMs) for reasoning-enhanced property prediction. Software updates, such as version 2025, have enhanced accuracy through refined training data and regulatory compliance features. Validation of these models relies on cross-validation techniques against experimental datasets to ensure reliability and generalizability. K-fold cross-validation, where data is partitioned into training and test folds iteratively, commonly yields R² values exceeding 0.85 for well-trained neural networks and QSAR models on benchmarks like PHYSPROP. External validation against independent experimental logP measurements confirms predictive power, with errors typically under 0.5 log units for in-distribution compounds, highlighting the robustness of these computational approaches. One common method to estimate the octanol-water partition coefficient () from aqueous solubility () relies on the general solubility equation () developed by Yalkowsky and colleagues for non-electrolytes. The GSE expresses (in mol/L) as: \log S_w = 0.5 - 0.01(T_m - 25) - \log P where T_m is the melting point in °C. Rearranging this equation yields an estimate for : \log P = 0.5 - \log S_w - 0.01(T_m - 25) This derivation assumes ideal solution behavior and is applicable primarily to neutral organic non-electrolytes with melting points below 200°C, where experimental solubility data is available but measurement is challenging due to low water solubility. The equation performs well for rigid molecules like polycyclic aromatics, with root mean square errors around 0.6 log units across diverse datasets, but accuracy decreases for flexible or polar compounds due to non-ideal activity coefficients. For supercooled liquids (T_m approximated as 25°C), the melting point term vanishes, simplifying to ≈ 0.5 - . When T_m is unknown, approximations substituting molecular weight () for the correction term have been explored in limited contexts, though they introduce additional error for solids. The distribution coefficient (log D) at a specific pH can be derived from log P and acid dissociation constant (pK_a) values using speciation fractions to account for ionization effects. For a monoprotic base, the neutral fraction f_n is given by: f_n = \frac{1}{1 + 10^{pH - pK_a}} Thus, log D = log P + log f_n. For monoprotic acids, f_n = 1 / (1 + 10^{pK_a - pH}), and log D follows analogously. For polyprotic or amphoteric compounds, the total neutral fraction is the sum of all neutral species fractions, often requiring iterative numerical solutions to handle coupled equilibria, especially when pK_a values are close. This approach assumes negligible ion partitioning into and is implemented in software like or , which automate the iteration for multi-site ionization and provide uncertainty estimates based on pK_a confidence intervals. The method is essential for ionizable drugs, where log D at physiological pH (e.g., 7.4) better predicts membrane permeability than log P. Other derivations of log P employ linear regressions from related physicochemical properties, such as boiling point (BP) or critical micelle concentration (). For homologous series like alkanes or alkylbenzenes, empirical relations like log P = a · (BP / 100) + b have been fitted, with coefficients a ≈ 0.04–0.06 and b ≈ -1.5 to -2.0, reflecting increased hydrophobicity with higher BP; these achieve r² > 0.9 for class-specific predictions but fail across diverse structures due to varying polar contributions. Similarly, for , CMC correlates inversely with log P via log CMC = c - d · log P (d ≈ 0.3–0.5), allowing rearrangement to log P = (c - log CMC) / d; this is useful for amphiphilic compounds where direct log P measurement is confounded by micellization, with applicability limited to ionic or nonionic series. These derivations assume solution ideality and neglect specific solute-solvent interactions, leading to systematic errors (up to 1–2 units) for amphiphiles, highly polar solutes, or compounds forming aggregates. For instance, hydrogen-bonding groups may overestimate P from due to unaccounted hydration energies. Recent refinements post-2018 incorporate parameters (HSP: δ_d for , δ_p for polar, δ_h for bonding) to better capture phase-specific interactions. By estimating activity coefficients in and octanol via the relation γ = (V / 2.303 RT) · Σ F_i (δ_i^{solvent} - δ_i^{solute})^2 (where F_i are interaction factors, V is ), P is derived as - (γ_water / γ_octanol); machine learning-enhanced HSP models improve accuracy for diverse datasets, reducing errors by 20–30% over classical GSE for polar organics.

Applications

Pharmacology and Drug Design

In and , the partition coefficient, particularly log P (the logarithm of the ), plays a pivotal role in by influencing drug and distribution. For oral , stipulates that compounds with log P greater than 5 are likely to exhibit poor due to excessive , which hinders and across gastrointestinal membranes. This guideline, derived from analysis of successful oral drugs, helps prioritize candidates during lead optimization to ensure adequate solubility and permeability. In distribution, log P is critical for crossing biological barriers, such as the (BBB), where a quantitative structure-activity relationship (QSAR) model predicts brain penetration via the equation: \log BB = 0.152 \cdot \log P - 0.0148 \cdot PSA + 0.139 Here, log BB represents the logarithm of the brain-to-blood concentration ratio, PSA is the polar surface area in Ų, and higher log P values (typically 1–3) enhance passive diffusion into the central nervous system (CNS) while avoiding efflux by transporters like P-glycoprotein. In pharmacodynamics, log P modulates receptor binding affinity and efficacy, especially for CNS-targeted drugs, where an optimal range of 2–3 balances lipophilicity for BBB penetration with minimal non-specific binding. This range ensures effective interaction with hydrophobic pockets in targets like dopamine or serotonin receptors, as deviations can reduce potency or increase off-target effects. For instance, in statin design, lipophilicity (log P) was optimized to achieve hepatic selectivity; hydrophilic statins like pravastatin (log P ≈ 0.2) limit extrahepatic distribution and reduce myotoxicity, while lipophilic ones like simvastatin (log P ≈ 4.7) enhance pleiotropic effects but require careful titration to avoid accumulation in non-target tissues. Similarly, in antipsychotics, log P tuning in atypical agents like aripiprazole (log P ≈ 4.5) improves CNS delivery and reduces extrapyramidal side effects compared to high-log P typical antipsychotics, by facilitating controlled diffusion and receptor occupancy. For , , , , and (ADMET) predictions, elevated log P (>4) correlates with increased risk of hERG inhibition, a major cause of QT prolongation and , as lipophilic compounds more readily access the channel's hydrophobic . QSAR models incorporating log P as a descriptor have quantified this, showing that reducing log P by 1 unit can increase hERG IC₅₀ (i.e., reduce potency of inhibition) by approximately 0.8 log units, guiding safer lead modifications. Addressing gaps in traditional models, updated QSAR approaches from the 2020s integrate log D (pH-dependent coefficient) to better predict outcomes in , where multiple drugs' interactions amplify variability and risks in patient populations. These models, often learning-enhanced, account for states at physiological to refine predictions for co-administered therapies.

Environmental and Agrochemical Sciences

In environmental sciences, the (Kow) serves as a key predictor of chemical in aquatic organisms, particularly through models that relate it to the bioconcentration factor (BCF). The U.S. Agency's (EPA) BCFBAF module in EPI Suite estimates BCF for non-ionizing organics using equations such as log BCF = 0.6598 log Kow - 0.333 for log Kow values between 1.0 and 7.0, adjusted by molecular corrections for factors like size and ; this approach highlights how higher log Kow values (typically >4) indicate greater potential for in fish lipids, influencing trophic transfer and ecological risk assessments. Similarly, soil-water partitioning is assessed via the organic carbon-water partition coefficient (Koc), which correlates strongly with Kow; empirical relationships, such as log Koc = 0.81 log Kow + 0.10 derived from large datasets of neutral organics, enable predictions of chemical sorption to , thereby estimating mobility and potential. In applications, partition coefficients guide the evaluation of fate in agricultural systems, particularly risks. Compounds with log P (or log Kow) >3 exhibit strong retention due to enhanced adsorption to , reducing downward migration and runoff into water bodies; this threshold is used in screening models like the Groundwater Ubiquity Score (GUS) index, where high log P values contribute to low indices (<1.8), promoting safer selection for minimizing environmental exposure. For instance, DDT's high log Kow of 6.91 drives its persistence in s and sediments, facilitating long-term bioaccumulation in food chains and contributing to its classification as a persistent organic pollutant under the Stockholm Convention. Regulatory frameworks incorporate partition coefficients for risk assessment in both domains. Under the European Union's REACH regulation, log Kow ≥3 triggers evaluation for bioaccumulative potential (B criterion) if BCF ≥2000, informing prioritization of testing and restrictions for substances posing ecosystem threats; this applies to agrochemicals, where it assesses chronic exposure via soil and water pathways. In contrast, neonicotinoids like imidacloprid demonstrate favorable plant uptake due to log D values (e.g., 0.57 at pH 7), reflecting their moderately lipophilic nature that enhances systemic translocation from roots to foliage while limiting non-target soil persistence. Emerging research addresses how climate change influences partitioning behaviors. Rising temperatures can alter log Kow by 10-20% over 20-30°C ranges for many organics, with the direction depending on the compound; for chlorobenzenes, it increases due to changes in phase interactions, potentially reducing chemical mobility in some ecosystems but enhancing volatilization; models project that 2-3°C global increases could elevate contaminant concentrations in surface waters by up to 50% for hydrophobic compounds, complicating agrochemical management and environmental monitoring.

Industrial and Material Applications

In the formulation of consumer products such as cosmetics, the logarithm of the octanol-water partition coefficient (log P) serves as a key parameter for assessing emulsion stability, particularly for oil-in-water systems where emollients and oils with log P values typically ranging from 4 to 6 exhibit optimal partitioning to prevent phase separation and enhance product shelf life. For instance, carrier oils influencing the partitioning of lipophilic compounds like cannabidiol (with log P ≈ 6.33) have been shown to improve interfacial tension and zeta potential, thereby boosting emulsion stability against creaming. Similarly, fragrance ingredients in dermocosmetic emulsions rely on log P to predict their distribution between oil and aqueous phases, ensuring controlled release and sensory performance. In metallurgy, particularly hydrometallurgical processes, partition coefficients guide the selective extraction of metal ions from aqueous solutions into organic phases, enhancing recovery efficiency in solvent extraction operations. The distribution coefficient (K), akin to the partition coefficient, quantifies the ratio of metal ion concentrations between phases, with values optimized through extractant selection like D2EHPA for scandium separation over other metals. In aqueous two-phase systems, such as those using PEG and salts, lead(II) ions achieve extraction efficiencies up to 74.4% at specific mass ratios, driven by favorable partitioning behaviors that minimize co-extraction of impurities. Coordination chemistry principles further underpin these processes, where ligand-metal interactions dictate partition selectivity in commercial flowsheets for base and precious metals. Partition coefficients play a crucial role in the food industry for retaining flavors and fragrances, influencing the equilibrium distribution of volatile compounds between the food matrix and headspace to maintain sensory profiles during processing and storage. In beverages and emulsions, partition coefficients for esters like ethyl butanoate vary with sucrose concentration, affecting release kinetics and perceived aroma intensity. For semi-solid matrices such as dairy products, these coefficients predict aroma retention in complex blends, with higher values indicating stronger binding to lipid phases and reduced volatilization. Octanol-air partition coefficients have proven effective in forecasting the release of nine volatile aromas, correlating with experimental headspace data to optimize formulation for consistent flavor delivery. In textile processing, dye partitioning relies on equilibrium partition coefficients to achieve uniform coloration and efficient uptake, especially for disperse dyes on hydrophobic fibers like polyethylene terephthalate (PET). The partition ratio, defined as the dye concentration in the yarn versus the dyeing bath, governs substantivity, with higher values (e.g., >10 for certain azo dyes) ensuring deep penetration and fixation. In aqueous two-phase systems for extraction from effluents, partition coefficients exceeding 26 for reactive dyes like Remazol Brilliant Blue R favor polymer-rich phases, aiding while recovering dyes for . Hydrophobic disperse dyes exhibit log P-driven partitioning that correlates with environmental persistence, informing sustainable practices. Solvent selection in paints and coatings often incorporates log P matching to ensure between binders, pigments, and , promoting even dispersion and film formation without defects like . Knowledge-based models using log P as a hydrophobicity descriptor facilitate of volatile compounds (VOCs) with alternatives that maintain partitioning akin to traditional like (log P ≈ 2.7). This approach has been integrated into guides, where log P guides the choice of bio-based to match rates and profiles in paints. Post-2020 trends emphasize sustainable alternatives to conventional solvents, with green solvents like bio-based options (e.g., dimethyl isosorbide) evaluated via partition coefficients to replicate performance in industrial applications while reducing environmental impact. These solvents, often with tunable log P values, enable greener extractions in and formulations, as seen in biobased replacements for dipolar aprotics that achieve comparable partitioning efficiencies. Such innovations address gaps in eco-friendly processing, prioritizing low-toxicity options with verified phase distribution properties.

Common Partition Systems

Octanol-Water System

The octanol-water partition system employs as the organic phase to model the lipophilic environment of biomembranes, owing to its amphiphilic structure featuring a nonpolar chain and a polar hydroxyl group that approximates the composition of cell membranes. serves as the hydrophilic aqueous phase, creating a biphasic setup that simulates solute distribution between biological fluids and lipid barriers. This choice of originated from evaluations in the 1960s and 1970s, where it was selected over other solvents for its practical utility in correlating with . Standard measurements follow guidelines from the , which specify equilibration at 25°C using mutually saturated solutions of and to ensure consistent compositions. The partition coefficient, expressed as log P = log₁₀(K_{ow}), is calculated from the equilibrium concentrations of the solute in each , with the OECD slow-stirring method recommended for compounds with log P up to 8.2 to minimize formation. High purity of , at least 99% and preferably purified by and , is essential, as impurities can modify solvent polarity and solute activity coefficients, leading to systematic errors in log P values. Representative experimental log P values from this system span a wide range, illustrating hydrophilicity (negative values) to (positive values greater than 3). The following table provides examples for common organic compounds, based on critically evaluated measurements at 25°C from the cited source (values for other compounds like (-0.07), (0.90), and aspirin (1.19) are standard literature averages from Hansch et al., 1995, with similar uncertainties):
Compoundlog P
-0.74
-0.31
Acetone-0.24
Phenol1.50
1.97
2.13
2.73
3.35
These values are recommended averages with uncertainties varying from ±0.03 to ±0.10, derived from direct experimental techniques like shake-flask methods. The octanol-water log P dataset underpins major physicochemical databases, such as those in the EPA's EPI Suite, enabling quantitative structure-activity relationship (QSAR) models for predicting environmental fate and toxicity.

Alternative Solvent Systems

Alternative solvent systems to octanol-water provide tailored assessments of solute partitioning by emphasizing specific molecular interactions, such as hydrophobic effects without hydrogen bonding or biomimetic environments that mimic biological interfaces. These systems are selected based on the target application, offering advantages in predicting behaviors like membrane permeability or environmental fate where octanol-water may overestimate polar interactions. Alkane-water systems, such as -water or cyclohexane-water, are widely used to model non-polar regions of membranes and evaluate passive permeability, as alkanes lack hydrogen-bonding sites that could confound results in more polar solvents. In these systems, log P values for solutes are typically 1-2 units lower than in octanol-water due to reduced of polar groups, providing a better with transcellular permeability in bilayers (r = 0.95 for ). For instance, high-throughput assays using barriers in parallel artificial permeability assays (PAMPA) derive intrinsic permeability from alkane-water log P, aiding early by focusing on hydrophobic driving forces. Cyclohexane-water partitioning similarly highlights apolar solute behavior, with experimental distribution coefficients measured via shake-flask methods to validate computational predictions of insertion. Biomimetic systems, including phospholipid-water or phosphatidylcholine-cholesterol bilayers, enhance biological relevance by replicating membrane compositions, allowing direct estimation of partitioning into cellular environments. These coefficients (Klipw) outperform octanol-water for predicting and , as they account for electrostatic and van der Waals interactions in lipid headgroups and tails. For example, partitioning into porcine lipid extract in PAMPA models correlates strongly (r² = 0.93) with blood- barrier permeability, capturing the nuanced effects of on . Such systems are particularly valuable for ionizable compounds, where pH-dependent partitioning reflects real physiological conditions. Chloroform-water partitioning serves as a specialized tool for studying hydrogen-bonding contributions to , as chloroform's weak hydrogen-bond acidity probes solute basicity without strong competition from the organic phase. Calculations for bases, such as 9-methyladenine, yield chloroform-water log P values that isolate hydrogen-bonding effects, revealing how intramolecular bonds alter partitioning by 0.5-1.5 log units compared to non-hydrogen-bonding solvents. This system has been applied to and derivatives to quantify intramolecular hydrogen bonding, aiding structure-activity relationships in . Vegetable oil-water systems are employed in to assess flavor retention and distribution in emulsions, where partitioning reflects real-world oil-water interfaces in processed foods. For and , bulk vegetable oil-water log K_ow values are -1.32 and 2.97, respectively (at 25°C), indicating hydrophilic partitioning for and lipophilic for , which influences release rates in oil-in-water emulsions. partitioning in edible oil-water binaries follows linear free energy relationships, with coefficients decreasing for hydrogen-bond donors due to oil's low , guiding formulation stability. Emerging ionic liquid-water systems support applications by enabling sustainable extractions and separations, with partition coefficients tuned via anion structure to minimize volatility and toxicity. Recent models predict gas-ionic liquid partitioning influenced by anion hydrophobicity, achieving errors below 0.5 log units for solutes, which extends to liquid-liquid systems for eco-friendly partitioning in pharmaceutical processing. These systems offer tunable interactions, such as enhanced for polar analytes, aligning with advancements in non-volatile solvents.

References

  1. [1]
    Partition Coefficient - an overview | ScienceDirect Topics
    A partition coefficient is the ratio of the concentration of a substance in one medium or phase (C1) to the concentration in a second phase (C2) when the two ...
  2. [2]
    Prediction of logarithm of n-octanol-water partition coefficient (logP ...
    Feb 23, 2023 · The logarithm of n-octanol-water partition coefficient (logP) is frequently used as an indicator of lipophilicity in drug discovery, ...
  3. [3]
    The Rule of Five for Non-Oral Routes of Drug Delivery - NIH
    The Rule of Five states that poor absorption or permeation is expected when MW>500, NHD>5, NHA>10 or log P>5. Although oral delivery remains dominant, other ...<|control11|><|separator|>
  4. [4]
    LogP vs LogD - What is the Difference? - ACD/Labs
    Jul 11, 2024 · LogP and logD are important values to consider during drug design as both give insight to the lipophilicity and hydrophilicity of compounds.
  5. [5]
    Partition Coefficient - an overview | ScienceDirect Topics
    Partition coefficients are defined as the concentration ratio of a chemical between two media at equilibrium (see also Chapter 1.04). The media can be gases ...<|control11|><|separator|>
  6. [6]
    Hermann Walther Nernst | Encyclopedia.com
    The case that Nernst studied theoretically and experimentally was the distribution of benzoic acid between water (phase w) and benzene (phase b).
  7. [7]
    Walther Nernst Memorial
    Nernst Distribution Law (1891). If two liquids (or solids) a and b are partially immiscible and if there is a third component i present in both phases which ...
  8. [8]
    Experimental Determination of Octanol–Water Partition Coefficients ...
    Feb 25, 2020 · Generally, neutral species showed higher affinity for the octanol phase than their respective (partly) ionized counterparts (Table 1). However, ...<|separator|>
  9. [9]
    Practical Understanding of Partition Coefficients | LCGC International
    Mar 1, 2023 · The partition coefficient, designated P (or, more commonly, log P), is a ratio of the concentration of a compound at equilibrium when disbursed ...
  10. [10]
    IUPAC - distribution ratio (D01817)
    ### Summary of Distribution Ratio (D01817) from IUPAC Gold Book
  11. [11]
    partition coefficient (P04437) - IUPAC
    This term is not recommended and should not be used as a synonym for partition constant, partition ratio or distribution ratio.Missing: definition | Show results with:definition
  12. [12]
  13. [13]
    [PDF] LogP—Making Sense of the Value - ACD/Labs
    According to 'Lipinski's Rule of 5' (developed at Pfizer) the logP of a compound intended for oral administration should be <5. A more lipophilic compound: ...Missing: five | Show results with:five
  14. [14]
    n-Octanol/Water Partition Coefficient (Kow/logKow)
    Jan 13, 2016 · n-Octanol/Water Partition Coefficient (Kow) is defined as the ratio of the concentration of a chemical in n-octanol and water at equilibrium at a specified ...
  15. [15]
    LogP and logD calculations - Chemaxon Docs
    The standard partition coefficient of ionized and unionized species calculated from the molecular structure is based largely on the atomic log P increments.
  16. [16]
    High-Throughput HPLC Method for the Measurement of Octanol ...
    Jun 5, 2025 · For measurements taken at a pH at which the compound is neutral, logD is equal to logP. However, in order to refine minor discrepancies caused ...
  17. [17]
    p-σ-π Analysis. A Method for the Correlation of Biological Activity ...
    Brief Article April 1, 1964. p-σ-π Analysis ... Smart citations by scite.ai include citation statements extracted from the full text of the citing article.
  18. [18]
    Hansch analysis 50 years on - Martin - Wiley Interdisciplinary Reviews
    Apr 18, 2012 · The original Hansch–Fujita QSAR continues to be performed to this day, fifty years since its inception. In addition, it has inspired vigorous research.
  19. [19]
    [PDF] Lipophilicity Descriptors: Understanding When to Use LogP & LogD
    For ionizable solutes, the compound may exist as a variety of different species in each phase at any given pH. D, the distribution coefficient, is the ...
  20. [20]
    Automated High Throughput pKa and Distribution Coefficient ... - NIH
    The distribution coefficient (logD) is defined as the ratio of the sum of concentrations of both charged and neutral species in the organic and aqueous phases.
  21. [21]
    Prediction of pH-Dependent Hydrophobic Profiles of Small ...
    Sep 28, 2017 · Because the distribution coefficient takes into account the partition of both neutral and ionic species of ionizable compounds, it provides an ...
  22. [22]
    Chapter 1: Physicochemical Properties - Books
    Feb 3, 2023 · The distribution coefficient, D, is the partition of a compound ... log D = log P − log(1 + 10pKa−pH). Equation 1.3. A theoretical plot ...
  23. [23]
    Assessing the Environmental Partitioning of Organic Acid Compounds
    At low pH values (approximately pH 1 to 2) , the neutral species is dominant in both phases, and the value of log D is Independent of pH. At intermediate pH ...
  24. [24]
    The Use of Molecular Descriptors To Model Pharmaceutical Uptake ...
    Dec 27, 2018 · The two descriptors log Kow and log D are both measures of hydrophobicity, but log D takes into account both neutral and ionizable species at a ...
  25. [25]
    [PDF] Nernst distribution law - BP Chaliha College
    Nernst (1891) studied the distribution of several solutes between different appropriate pairs of solvents. He gave a generalization which governs the ...Missing: original | Show results with:original
  26. [26]
    The n-octanol and n-hexane/water partition coefficient of ...
    ... partition coefficients of the same substances in different distribution systems. ... Nernst's distribution law (Nernst, 1891) and is called the `partition ...
  27. [27]
    Simple method to calculate octanol–water partition coefficient of ...
    The first method to evaluate of logP was proposed by Fujita, Iwasa, and Hansch (1964). In their approach, they considered logP to be an additive ...Missing: original | Show results with:original
  28. [28]
    Using Thermodynamics and Simulations to Understand Selectivity ...
    Feb 1, 2023 · Equation 4 is the classical Gibbs' equation, which relates the partition coefficient and the standard free energy change. As seen from ...
  29. [29]
    Thermodynamics of Water-octanol and Water-cyclohexane ... - MDPI
    This reduces polarity and intermolecular hydrogen bonding ability, and might therefore be expected to increase partition coefficient.<|separator|>
  30. [30]
    [PDF] the thermodynamics of partitioning of phenolic compounds between ...
    The low positive enthalpy of transfer for p-nitro phenol reflects the negative enthalpy contribution due to hydrogen bonding of the hydroxyl group with the ...
  31. [31]
    [PDF] Partition coefficients and their uses
    Feb 27, 2020 · ... distribution of a solute between two phases in which it is soluble ... ratio of the concentrations of solute distributed between two ...
  32. [32]
    Evaluating the Salting-Out Effect on the Organic Carbon/Water ...
    Jul 11, 2016 · We present new measurements of the salting-out constants (K s ) for partition ratios between water and organic carbon (K OC ) and between water and dissolved ...
  33. [33]
    Three- and Multi-Phase Extraction as a Tool for the Implementation ...
    Sep 25, 2022 · The three-phase multi-stage extraction is carried out in a cascade of bulk liquid membrane separation stages, each comprising two interconnected (extraction ...
  34. [34]
    Influence of emulsifier concentration on partition behavior and ...
    May 1, 2019 · In emulsions, the API partitions between both phases according to the partition coefficient and may insufficiently be protected from degradation ...Missing: multi- | Show results with:multi-
  35. [35]
    Empirical relationships between the 1-octanol/water partition ...
    Empirical relationships of aqueous solubility and octanol‐water partition coefficient of hydrocarbons with different molecular descriptors. Toxicological ...
  36. [36]
    Prediction of bioconcentration potential of organic compounds using ...
    Prediction of bioconcentration potential of organic compounds using partition coefficients derived from reversed phase thin layer chromatography.
  37. [37]
    [PDF] Poster Text: In Silico Prediction of Physicochemical Properties of ...
    Abbreviations: log BCF = log of bioconcentration factor; log P = octanol–water partition coefficient; log S = water solubility; log VP = log of vapor pressure.
  38. [38]
    Chapter: 5 Physicochemical Properties and Environmental Fate
    This equation uses regression-derived correlation with logP and melting point (MP) for solids: logS = 0.8 – logP – 0.01(MP – 25). Other factors that ...
  39. [39]
    Physicochemical Properties of Higher Nonaromatic Hydrocarbons
    Literature data on log Kow for n-alkanes expressed as a function of the molar volume. Lines represent linear regression: C1 to C8 n-alkanes. 共Refs. 23 and ...
  40. [40]
  41. [41]
    [PDF] Partition coefficient of acetic acid
    Feb 17, 2020 · Fill your burette with the NaOH titrant. 2. Add the assigned volumes of DI water and octanol to a separatory funnel. Make sure the stopcock is ...
  42. [42]
  43. [43]
    Application of a modified LSER model to retention on a ... - NIH
    The modified LSER model including the P solute descriptor was found to be viable for analyte sets with both neutral and ionizable compounds [38]. It was noted, ...
  44. [44]
    Chromatographic Approaches for Measuring Log P - ResearchGate
    retention parameters were improved by separating compounds in two classes, i.e. the correlation between log k obtained on the LC - ABZ column and log P [13] . ...
  45. [45]
    [PDF] assessment of reverse - phase - ECETOC
    In order to correlate the measured HPLC or TLC data of a compound with its P, a calibration graph of log P vs chromatographic data is needed for a number ( ...
  46. [46]
    PH-Metric log P. 4. Comparison of Partition Coefficients Determined ...
    Values of log P obtained both by potentiometry and by partition HPLC, which ranged from 0.3 to 5.4, were in very good acordance with literature values.
  47. [47]
    pH-metric log P. 4. Comparison of partition coefficients determined ...
    Values of log P obtained both by potentiometry and by partition HPLC, which ranged from 0.3 to 5.4, were in very good acordance with literature values.
  48. [48]
    Recent advances in lipophilicity measurement by reversed-phase ...
    It offers several practical advantages, including speed, reproducibility, insensitivity to impurities or degradation products, broad dynamic range, on-line ...Missing: limitations | Show results with:limitations
  49. [49]
    Rapid Determination of Lipophilicity: Establishment and Application ...
    Nov 30, 2023 · The two RP-HPLC methods (Method 1 and Method 2) and the shake-flask method enable accurate and efficient log P determination for compounds at ...
  50. [50]
    High-throughput logPo/w determination from UHPLC measurements
    Aug 5, 2016 · The proposed equations allow an accurate determination of logPo/w with standard errors in the range of 0.4 units. Keywords: CHI; Chromatographic ...
  51. [51]
    Electrochemical Determination of Partition Coefficients of Drugs
    The method involves cyclic voltammetry at the polarizable interface between two immiscible electrolyte solutions.Missing: partitioning | Show results with:partitioning
  52. [52]
    Hydrodynamic Voltammetry at the Liquid/Liquid Interface:  The Rotating Diffusion Cell
    ### Summary of Hydrodynamic Voltammetry at ITIES and Relation to Partitioning/Partition Coefficient
  53. [53]
    Novel electrochemical approach to the determination of the partition ...
    Aug 7, 2025 · A novel electrochemical approach to the determination of the partition coefficient of neutral weak bases is presented.
  54. [54]
    Determination of octanol–water partition coefficients using a micro ...
    Aug 5, 2025 · This method was used to accurately and rapidly determine octanol-water partition coefficients (Kow), determining identical Kow as the shake- ...<|control11|><|separator|>
  55. [55]
    Microfluidic impedance spectroscopy as a tool for quantitative ... - NIH
    Jul 13, 2012 · A microfluidic device that is able to perform dielectric spectroscopy is developed. The device consists of a measurement chamber that is 250 μm ...Missing: partition | Show results with:partition
  56. [56]
    Quantification of Small Molecule Partitioning in a Biomolecular ...
    Jul 24, 2025 · Several compounds give multiple NMR signals, resulting in multiple PCs for these compounds. As each compound only has one partition coefficient, ...1 Introduction · 4 Experimental Section · Small Molecule Nmr Samples
  57. [57]
    Octanol–Water Partition Coefficient Measurement by a Simple 1H ...
    Sep 28, 2017 · We describe a simple miniature shake-flask method to measure the octanol–water partition coefficient of an organic compound.Introduction · Method Basics · Representative Procedure · Discussion
  58. [58]
    Advantages of an Electrochemical Method Compared to the ... - NIH
    Electrochemical techniques have their advantages because of their simplicity, low cost and speed. The only condition for this method of monitoring enzyme ...
  59. [59]
    Troubleshooting the influence of trace chemical impurities on ...
    Apr 25, 2024 · This work highlights the utility of real-time electrochemical potential measurements as a tool for benchmarking of nanoparticle syntheses and troubleshooting ...
  60. [60]
    AI-Enhanced Electrochemical Sensing Systems: A Paradigm Shift ...
    Aug 28, 2025 · This review examines the integration of AI technologies, particularly machine learning and deep learning algorithms, in enhancing sensor design, ...Missing: post- | Show results with:post-
  61. [61]
    AI-Empowered Electrochemical Sensors for Biomedical Applications
    AI algorithms can further improve the accuracy, sensitivity, and repeatability of electrochemical sensors through the screening and performance prediction of ...Missing: partition post-
  62. [62]
    Machine‐Learning‐Aided Advanced Electrochemical Biosensors
    Jun 9, 2025 · This review highlights recent ML applications in advanced and nanomaterial-enhanced electrochemical biosensing, categorized into biocatalytic ...
  63. [63]
    Exploring the octanol–water partition coefficient dataset using deep ...
    Jun 14, 2021 · We have selected the octanol-water partition coefficient (log P) as an example, which plays an essential role in environmental chemistry and ...
  64. [64]
    Pushing the limit of logP prediction accuracy: ChemAxon's results for ...
    Chemaxon's logP method achieved the highest accuracy with 0.31 RMSE, outperforming references, and showed high accuracy on never-seen molecules.Missing: Schrödinger neural networks
  65. [65]
    Review of the latest progress of AI and Machine Learning methods ...
    Oct 6, 2025 · Schrödinger's Jaguar, DFT-based software, often achieved the highest accuracy in prediction. ... neural networks to perform reaction prediction ...
  66. [66]
    [PDF] Review of QSAR Models and Software Tools for predicting ...
    Such methods include 3D-QSAR (e.g. Comparative Molecular Field Analysis, CoMFA); quantitative molecular similarity analysis (QMSA), based on experimental data ...
  67. [67]
    QSAR application for the prediction of compound permeability with ...
    An in silico model was recently developed for the prediction of parallel artificial membrane permeability assay (PAMPA) permeability using logP, pKa, and PSA ...
  68. [68]
    Universal Quantum Mechanical Model for Solvation Free Energies ...
    We present a new solvation model for predicting free energies of transfer of organic solutes from the gas phase to aqueous and organic solvents.Missing: logP derivation
  69. [69]
    LogP prediction performance with the SMD solvation model and the ...
    The model calculates solvation free energy differences using the M06-2X functional with SMD implicit solvation and the def2-SVP basis set. This model was ...Missing: derivation | Show results with:derivation
  70. [70]
    Improving solvation energy predictions using the SMD solvation ...
    COSMO is used with AM1, PM3, and PM6 as implemented in the MOPAC program.28 The null model is constructed by setting all the predicted solvation free energies ...Missing: logP derivation
  71. [71]
    Improved Lipophilicity and Aqueous Solubility Prediction with ... - MDPI
    We argue that combining models with different key aspects help make graph neural networks deeper and simultaneously increase their predictive power.
  72. [72]
    Multi-fidelity graph neural networks for predicting toluene/water ...
    Aug 8, 2024 · This paper uses multi-fidelity learning with graph neural networks to predict toluene/water partition coefficients, achieving improved accuracy ...Missing: 2020s | Show results with:2020s
  73. [73]
    Dimensionally reduced machine learning model for predicting single ...
    Jan 19, 2023 · MF-LOGP, a new method for determining a single component octanol–water partition coefficients ( L o g P ) is presented which uses molecular ...
  74. [74]
    Prediction of n-Octanol/Water Partition Coefficients from PHYSPROP ...
    Prediction of n-Octanol/Water Partition Coefficients from PHYSPROP Database Using Artificial Neural Networks and E-State Indices. Click to copy article link ...
  75. [75]
    Experimental and computational approaches to estimate solubility ...
    In the discovery setting 'the rule of 5' predicts that poor absorption or permeation is more likely when there are more than 5 Hbond donors, 10 Hbond ...
  76. [76]
    Rapid calculation of polar molecular surface area and its application ...
    A method for the rapid computation of polar molecular surface area (PSA) is described. It is shown that consideration of only a single conformer when computing ...Missing: log BB logP
  77. [77]
    Medicinal chemical properties of successful central nervous system ...
    By far, the most important is hydrogen bonding ability. These early studies underlined the importance of LogP in understanding the activity of CNS drugs and ...
  78. [78]
    Classics in Chemical Neuroscience: Aripiprazole - PMC
    Neuroimaging studies demonstrate that clinical improvement in positive symptoms using first-generation antipsychotics ... Log P value of 4.55. These ...
  79. [79]
    The Role of Structure and Biophysical Properties in the Pleiotropic ...
    Nov 19, 2020 · The most widely used method of communicating statin lipophilicity is through the logP partition coefficient (typically water and n-octanol) for ...
  80. [80]
    hERG toxicity assessment: Useful guidelines for drug design
    Jun 1, 2020 · QT prolongation and hERG channel interactions are surrogate markers of cardiotoxicity. Drug which induces QT interval prolongation has been ...
  81. [81]
    Ligand-based prediction of hERG-mediated cardiotoxicity ... - Frontiers
    Sep 4, 2022 · Cardiotoxicity is a common side effect of drugs, and one of the causes for it is the off-target interaction with different voltage-gated ion ...
  82. [82]
    EPI Suite™-Estimation Program Interface | US EPA
    Aug 25, 2025 · KOWWIN™: Estimates the log octanol-water partition coefficient, log KOW, of chemicals using an atom/fragment contribution method. AOPWIN ...Missing: formula | Show results with:formula
  83. [83]
    Estimating Koc for persistent organic pollutants - ScienceDirect.com
    The n-octanol/water partition coefficient (Kow) is commonly used to predict the soil or aquatic particle water partition coefficient normalized to organic ...
  84. [84]
    NAFTA Guidance Document for Conducting Terrestrial Field ... - EPA
    The basic TFD study focuses on pesticide dissipation from the soil surface layer in a bareground study; it can be used to estimate field degradation only when ...<|control11|><|separator|>
  85. [85]
    Environmental Impact of Pesticides: Toxicity, Bioaccumulation and ...
    Jul 10, 2025 · The lipophilic nature of DDT (log Kow = 6.91) and its resistance to metabolic degradation resulted in extensive bioaccumulation in fatty tissues ...
  86. [86]
    Guidance on Information Requirements and Chemical Safety ...
    This guidance describes the information requirements under REACH with regard to substance properties, exposure, use and risk management measures, in the context ...Missing: log Kow
  87. [87]
    Uptake and translocation of pesticides in pepper and tomato plants
    Nov 29, 2024 · This can be explained by the low apparent octanol to water partition coefficient, log D, of this compound (0.95, Supporting Information, Table ...
  88. [88]
    Temperature Dependence of Octanol−Water Partition Coefficient ...
    The van't Hoff plots of log KOW versus T-1 exhibit linearity with values of KOW increasing by 10%−14% over this temperature range. The enthalpy of phase change ...Missing: Kow | Show results with:Kow
  89. [89]
    INFLUENCE OF GLOBAL CLIMATE CHANGE ON CHEMICAL FATE ...
    We focus on the 2 and 3°C increases, which represent the fastest growth and highest temperature scenarios, respectively (Fig. 4C and D). Concentration ratios >1 ...Missing: Kow | Show results with:Kow
  90. [90]
    [PDF] Influence of the emollient structure on the properties of cosmetic ...
    Oct 25, 2019 · We tried to classify the selected emollients by calculating logP. The octanol-water partition coefficient (KO-W) is the ratio of a ...
  91. [91]
    Impact of carrier oil on interfacial properties, CBD partition and ...
    Oct 1, 2022 · CBD molecule has a high partition coefficient value (log P = 6.33 ... emulsion properties, CBD partition, and emulsion stability (Fig.2.8. 1. Creaming Index (ci) · 3. Results And Discussion · 3.4. Zeta Potential
  92. [92]
    Fragrance in dermocosmetic emulsions: From microstructure to skin ...
    Aug 25, 2023 · Partition coefficients often expressed in logarithmic form log P (Equation 2), are useful in estimating the affinity and distribution of ...
  93. [93]
    Recent Advances in Solvent Extraction for the Efficient Recovery of ...
    Jan 6, 2025 · They reported that using D2EHPA alone resulted in higher distribution ratio and better separation for Sc over other metals, and Sc was more ...
  94. [94]
    Extraction of Lead Ions and Partitioning Behavior in Aqueous ...
    A maximum lead(II) extraction efficiency of 74.4% was achieved using the PEG4000/(NH4)2SO4 system with a mass fraction ratio of PEG4000 to (NH4)2SO4 of 0.2:0.12 ...
  95. [95]
    the coordination chemistry behind extractive metallurgy
    Oct 3, 2013 · This tutorial review describes the coordination chemistry of commercial solvent extraction processes that are used to achieve the separation and concentration ...Oxime Extractants · Metalate Anion Extraction · Metal-Salt Extractants
  96. [96]
    Flavor Retention and Release from Beverages: A Kinetic and ... - NIH
    Nahon et al. reported partition coefficients for ethyl acetate, methyl butanoate, ethyl butanoate, hexanal, and octanal as a function of sucrose concentration, ...
  97. [97]
    Determination of partition coefficients in complex semi-solid matrices
    This study enabled us to determine the partition coefficient of a blend of aroma compounds in a complex food matrix by way of the PRV method. The results were ...Missing: fragrance | Show results with:fragrance
  98. [98]
    Effectiveness of octanol-air partition coefficients in predicting aroma ...
    Jan 15, 2025 · This study examined the relationship between three partition coefficients (water-air, octanol-water, and octanol-air) and the release of nine volatile aromas ...
  99. [99]
    Dye uptake and partition ratio of disperse dyes between a PET yarn ...
    The equilibrium partition coefficient of the dye is the ratio between the equilibrium concentration of the dye in the yarn and that in the dyeing medium, and is ...
  100. [100]
    Thermodynamics of Textile Dyes Partitioning in Alcohol‐Based ...
    May 27, 2025 · The partition of a textile dye, Remazol Brilliant Blue R, was evaluated and presented partition coefficients (K) higher than 26.4 and ...
  101. [101]
    i. solubility and partitioning of some hydrophobic dyes and
    This study assesses the environmental behavior of about 50 dyes using solubilities, partition coefficients, and vapor pressures, focusing on disperse dyes and ...
  102. [102]
    Interactive Knowledge-Based Kernel PCA for Solvent Selection
    Mar 13, 2025 · ... solvent selection and substitution. The octanol: water partition coefficient (Log P), which describes the hydrophobicity of a solvent, is ...
  103. [103]
    Solvent Sustainability Guide for Paints & Coatings
    Jun 27, 2024 · A guide has been developed, highlighting various sustainability criteria of solvents used in the paints and coatings industry.
  104. [104]
    Dimethyl Isosorbide: An Innovative Bio-Renewable Solvent for ...
    Jun 24, 2025 · This study highlights DMI's promise as a sustainable bio-renewable alternative to conventional organic solvents used as eluents in LC, ...Missing: post- | Show results with:post-
  105. [105]
    Replacement of Less-Preferred Dipolar Aprotic and Ethereal ...
    Feb 24, 2022 · In both studies, less desirable polar aprotic solvents were successfully replaced with greener, safer, more sustainable alternatives.
  106. [106]
    Eco-Friendly Alternatives to Conventional Solvents: Innovations and ...
    Oct 12, 2025 · Bio-based solvents, one of the prominent types of green solvents, are sustainable substitutes for conventional petrochemical solvents. Bio-based ...
  107. [107]
    Quantifying molecular partition into model systems of biomembranes
    The octanol/water partition coefficient has been the subject of extensive ... Partitioning of substituted phenols in liposome–water, biomembrane–water, and ...
  108. [108]
    [PDF] Octanol-Water Partition Coefficients of Simple Organic Compounds
    The octanol-water partition coefficient of a substance X at a given temperature is, by general consent,' represented by P and defined by (for reasons explained ...
  109. [109]
    [PDF] Octanol-Water Partition Coefficients of Simple Organic Compounds
    Oct 15, 2009 · Octanol-water partition coefficients (log P) are for 611 simple organic compounds, representing all principal classes.
  110. [110]
    Partitioning of Solutes in Different Solvent Systems: The Contribution ...
    It is shown that the water-octanol partition coefficient is dominated by solute hydrogen-bond basicity, which favors water, and by solute size, which favors ...Missing: studies | Show results with:studies
  111. [111]
    Drug–Membrane Permeability across Chemical Space
    Jan 8, 2019 · These were in logarithmic units 0.5, -1.0, and 0 for octanol, hexadecane, and PGDP solvent systems, resp. Below these distribution coeffs ...
  112. [112]
    The PAMPA technique as a HTS tool for partition coefficients ...
    Partition coefficients in alkane–water system have already been derived from the intrinsic permeability coefficient (log P0) through hexadecane liquid membrane ...
  113. [113]
    Measuring experimental cyclohexane-water distribution coefficients ...
    In this work, we describe the experimental protocol we utilized for measurement of cyclohexane-water distribution coefficients, report the measured data,
  114. [114]
    Prediction of Phospholipid–Water Partition Coefficients of Ionic ...
    The partition coefficient of chemicals from water to phospholipid membrane, Klipw, is of central importance for various fields.
  115. [115]
    Partitioning into phosphatidylcholine–cholesterol membranes
    May 2, 2023 · The membrane–water coefficients overcome many of the challenges associated with octanol–water partitioning, are more biologically relevant, and ...
  116. [116]
    Physicochemical Selectivity of the BBB Microenvironment Governing ...
    The study uses a PAMPA model with porcine brain lipid extract to mimic the BBB and predict its permeability, showing a 93% variance prediction.
  117. [117]
    Calculation of Chloroform/Water Partition Coefficients for the N ...
    The chloroform/water partition coefficients have been determined for five N-methylated nucleic acid bases (9-methyladenine, 9-methylguanine, 1-methylcytosine, ...
  118. [118]
    The solvation, partitioning, hydrogen bonding, and dimerization of ...
    We present M06-2X density functional calculations of the chloroform/water partition coefficients of cytosine, thymine, uracil, adenine, and guanine.
  119. [119]
    Hydrogen-bonding abilities for phenols assessed by quantitative ...
    Hydrogen-bonding abilities for phenols assessed by quantitative analyses of their partition coefficients derived from different partitioning systems · Authors.Missing: studies | Show results with:studies
  120. [120]
    Partitioning of caffeine and quinine in oil‐in‐water emulsions and ...
    Nov 15, 2022 · The bulk vegetable oil–water partition coefficient of caffeine and quinine was determined by a shake-flask method as log Kow = −1.32 and ...
  121. [121]
    Partitioning of Antioxidants in Edible Oil–Water Binary Systems and ...
    The partition coefficient PWO is, thus, defined as the ratio of the effective concentrations of the solute in the oil and water phases (moles of antioxidant ...
  122. [122]
    Exploring the Influence of Ionic Liquid Anion Structure on Gas-Ionic ...
    This article presents an in-depth investigation into the influence of anionic structures of ionic liquids (ILs) on gas-ionic liquid partition coefficients ...Missing: 2020s | Show results with:2020s