Fact-checked by Grok 2 weeks ago

Partial charge

A partial charge, also known as an partial charge, refers to a non-integer (fractional) assigned to an individual within a , arising from the uneven distribution of electrons due to differences in , inductive effects, or other factors influencing . These charges are typically denoted using the Greek symbol, with δ⁺ indicating a partial positive charge on an that has lost and δ⁻ for a partial negative charge on an that has gained it, contrasting with full integer charges like those in ions. Unlike formal charges, which are integers based on electrons, partial charges provide a more nuanced representation of molecular ionicity and are not direct quantum mechanical observables but rather interpretive proxies derived from electron density distributions. In and computational modeling, partial charges play a pivotal role in elucidating chemical , molecular reactivity, and intermolecular interactions, such as dipole-dipole forces and , which govern properties like , boiling points, and . They are essential for parameterizations in simulations and , where accurate electrostatic potentials are required to predict molecular behavior. Historically, the concept traces back to Robert Mulliken's 1955 population analysis, which partitioned between atoms, though it has since evolved to address ambiguities in charge assignment. Various methods exist for calculating partial charges, broadly categorized into electrostatic potential fitting (e.g., CHELPG and RESP schemes), orbital-based partitioning (e.g., Mulliken and Natural Population Analysis), and partitioning (e.g., Bader's Quantum Theory of Atoms in Molecules and Hirshfeld methods). These approaches often yield differing values due to their underlying assumptions, highlighting the context-dependent nature of partial charges, yet they converge sufficiently for practical applications in chemistry and . Recent advances, such as ionic scattering factors modeling via 3D , enable experimental determination of partial charges at an absolute scale, validating computational predictions and extending their utility to crystalline materials like organic pharmaceuticals and zeolites.

Fundamentals

Definition

A partial charge is a non-integer, fractional charge assigned to an atom or molecular fragment, arising from the uneven of electrons within a or due to differences in atomic electronegativities. These charges are typically denoted using the Greek letter (δ), with δ⁺ indicating a partial positive charge and δ⁻ a partial negative charge; their magnitudes are usually between 0 and 1 in units of the elementary charge (e). For example, in the (HCl) molecule, the difference between (2.20) and (3.16) leads to chlorine bearing a partial negative charge of approximately -0.18 e and a partial positive charge of +0.18 e. In contrast to formal charges, which are integer values calculated by assuming equal sharing of bonding electrons between atoms, partial charges capture the quantum mechanical effects of electron delocalization, where electrons are not fully transferred but probabilistically distributed according to densities.

Physical Origin

Partial charges originate in polar covalent bonds, where electrons are shared unequally between atoms due to differences in , the tendency of an atom to attract shared electrons in a bond. On the Pauling electronegativity scale, which quantifies this property based on bond energy data, atoms like (4.0) exhibit high values compared to (2.20), leading to greater shifts and thus larger partial charges. This unequal sharing results in the more electronegative atom acquiring a partial negative charge (δ⁻) and the less electronegative atom a partial positive charge (δ⁺), without full as in ionic bonds. From a quantum mechanical perspective, partial charges arise from the delocalized nature of in , governed by the and described by molecular orbitals. In these orbitals, electrons are not confined to individual atoms but smear across the , creating non-integer charge distributions that reflect interactions without discrete transfers. This smearing is captured by the total electron density ρ(r), a fundamental quantum that partitions unevenly due to atomic differences, leading to effective charge separations observable in properties like dipole moments. Electrostatics provides another view, treating partial charges as effective point charges placed at atomic nuclei to approximate the molecule's electrostatic potential (MEP), the energy of interaction with a unit positive test charge. This model simplifies the continuous into discrete charges for computing intermolecular forces, with the partial charges chosen to reproduce the MEP on a molecular surface. A representative example is the water molecule (H₂O), where oxygen's (3.44) exceeds that of (2.20), causing the O–H bonds to be polar with δ⁻ on oxygen and δ⁺ on each , resulting in an overall molecular of about 1.85 D.

Determination Methods

Experimental Techniques

Experimental techniques for inferring partial charges rely on measuring observable molecular properties that arise from uneven charge distributions, such as electric moments and maps derived from data. These methods provide empirical insights into charge separation without direct computation, often requiring modeling to assign partial charges. One of the earliest and most direct approaches involves measuring molecular moments, which quantify the overall charge asymmetry in a . The \vec{\mu} is given by \vec{\mu} = \sum_i q_i \vec{r}_i, where q_i are partial charges on atoms and \vec{r}_i are their position vectors relative to the center of charge; experimental values thus imply average partial charges when combined with known geometries. experiments originated in the 1910s with Peter Debye's measurements of dielectric constants in polar gases, providing the first evidence for partial charges in molecules like and HCl. For instance, the measured of HCl is 1.08 D, corresponding to partial charges of approximately +0.18 e on H and -0.18 e on Cl, assuming a of 1.27 . Electron density mapping through diffraction techniques offers a more detailed view of charge distributions. X-ray crystallography determines the electron density \rho(\vec{r}) from structure factors obtained via Fourier transformation of intensities, allowing multipole modeling to refine partial charges by fitting aspherical electron distributions around . This approach has been applied to small molecules and crystals, revealing charge shifts due to bonding effects, as detailed in comprehensive reviews of charge-density analysis. Gamma-ray complements X-ray methods by using higher-energy photons that scatter more sensitively from positions while still probing electron density, enabling precise determination of charge distributions in materials with heavy or complex bonding; for molecular systems, it provides higher accuracy in multipole refinements for partial charge assignment. A recent advance in experimental determination involves electron diffraction on molecular crystals, which accesses partial charges through ionic scattering factors derived from the crystal structure. This method models as arising from ionic potentials, allowing absolute-scale assignment of partial charges that align with computational predictions. It is particularly useful for crystalline materials like pharmaceuticals and zeolites, providing validation for theoretical models. Spectroscopic methods indirectly infer partial charges through their influence on nuclear environments and vibrational modes. In (NMR) , chemical shifts correlate with local , as deshielding occurs when partial positive charge reduces shielding around a ; for example, 13C NMR shifts in substituted hydrocarbons show linear relationships with Mulliken or partial charges, enabling charge estimation from experimental spectra. () probes charge effects on bonding by measuring vibrational frequencies and intensities, where bond polarity—driven by partial charges—alters the force constant and dipole derivative, shifting frequencies for polar bonds like C-O (around 1000-1300 cm⁻¹) compared to nonpolar ones.

Computational Approaches

Computational approaches to partial charges derive atomic charge distributions from quantum mechanical wavefunctions or electron densities, providing predictive tools for molecular modeling without direct experimental measurement. These methods partition the total electron density among atoms using theoretical frameworks, often implemented in quantum chemistry software. They are broadly grouped by partitioning strategy, with early methods relying on basis set orbitals and later ones on density-based, topological, or electrostatic potential analyses. One of the earliest methods is the Mulliken population analysis, introduced in , which allocates by sharing overlap populations between basis functions centered on different atoms. The atomic charge q_A for atom A is computed as q_A = Z_A - \left[ \sum_{\mu \in A} P_{\mu\mu} + \sum_{\mu \in A} \sum_{\nu \notin A} P_{\mu\nu} S_{\nu\mu} \right], where Z_A is the nuclear charge, P is the , and S is the overlap matrix. This approach is computationally efficient but sensitive to basis set choice due to its reliance on overlap matrices. Related orbital-based methods include Löwdin charges, developed in 1950, which use symmetric orthogonalization of atomic orbitals to compute populations, yielding more stable charges that are less dependent on basis set superposition. Density-based methods partition the molecular using weight functions or iterative schemes. Hirshfeld charges, proposed in 1977, iteratively allocate based on weight functions from superimposed promolecular densities, ensuring charges reflect the molecular . Topology-based methods define atomic regions via properties of the . Bader's (AIM) theory, formalized in the 1990s, delineates atomic basins using zero-flux surfaces where the gradient is perpendicular to the surface, then integrates the over these basins for partial charges. This provides a physically rigorous, basis-set-independent partitioning but requires accurate from high-level methods. Electrostatic potential (ESP)-fitting methods determine charges by minimizing the difference between the molecular ESP (computed from ) and that reproduced by point charges on atoms, evaluated on a grid surrounding the . Common schemes include CHELPG (Charges from Electrostatic Potentials using a Grid), which fits to ESP on a fine grid while constraining to reproduce the , and RESP (Restrained Electrostatic Potential), which adds restraints to avoid and improve transferability. These methods are particularly valuable for force fields in simulations as they directly target intermolecular . Modern developments include charge partitioning methods optimized for diverse systems, such as periodic structures. The DDEC6 method, published in 2016, employs a delocalization-charge partitioning that uses fixed reference charges and iterative stockholder partitioning to yield transferable, chemically intuitive charges suitable for molecules, solids, and surfaces. Post-2016 extensions of the DDEC framework, building on earlier variants like DDEC3, enhance applicability to periodic systems by incorporating efficient s for band structure and surface calculations while maintaining even-tempered charge distributions. In the 2020s, techniques have emerged for partial charge assignment, particularly in development, where neural networks trained on quantum-derived densities predict charges rapidly for large datasets. For instance, neural networks can learn electrostatic potentials to assign charges that reproduce moments, offering scalability for biomolecular simulations. These approaches often validate against experimental moments derived from . Quantum chemistry software such as Gaussian and routinely implements these methods for wavefunction-derived charges, enabling users to compute Mulliken, Löwdin, Hirshfeld, , DDEC, CHELPG, and RESP charges from standard output files.

Applications

Molecular Modeling

In molecular modeling, partial charges play a central role in parameterizing non-bonded electrostatic interactions within empirical force fields, which are essential for simulating molecular behavior. These charges are typically fixed values assigned to atoms based on quantum mechanical calculations or experimental fitting, enabling the computation of electrostatic energies through . The electrostatic between two atoms i and j is given by: E = \frac{1}{4\pi\epsilon_0} \frac{q_i q_j}{r_{ij}} where q_i and q_j are the partial charges, r_{ij} is the interatomic distance, and \epsilon_0 is the vacuum permittivity. This formulation approximates the long-range Coulombic interactions in classical force fields like AMBER and CHARMM, which rely on such terms to model biomolecular systems accurately. Partial charges are integral to molecular dynamics (MD) and Monte Carlo (MC) simulations, where they contribute to the total potential energy function alongside bonded terms like bonds, angles, and dihedrals. In the AMBER force field, partial charges are derived using restrained electrostatic potential (RESP) fitting to ab initio electron densities, ensuring compatibility with simulations of proteins, nucleic acids, and small organic molecules. Similarly, the CHARMM force field employs partial atomic charges optimized against quantum mechanical data and experimental observables, facilitating biomolecular simulations that capture structural dynamics and thermodynamic properties. These fixed-charge models have been widely adopted for their computational efficiency in large-scale simulations. A key application of partial charges in molecular modeling is , where they enable predictions of solvation energies and -protein affinities by quantifying desolvation penalties and electrostatic contributions to complex formation. For instance, accurate partial charge assignments in force fields like CHARMM are critical for modeling the electrostatic interactions that influence , as deviations in charge distribution can significantly alter predicted energies. In recent advancements, polarizable force fields such as have incorporated dynamic partial charges that respond to environmental through inducible atomic dipoles and higher-order multipoles, improving simulations of charge fluctuations in solvated systems compared to fixed-charge approaches. This integration, developed starting in the early with major extensions in the , enhances the accuracy of MD simulations for flexible molecules in contexts. Recent developments as of 2025 include models for predicting partial atomic charges, particularly in complex systems like metal-organic frameworks (MOFs) and for long-range interactions in simulations. For example, models such as PACMOF2 predict density-derived electrostatic and wavefunction-derived charges to improve parameterization and property predictions in porous materials. Similarly, approaches have been applied to generate effective partial charges that capture conformer-dependent variations and enhance in biomolecular simulations. These methods offer greater efficiency and accuracy for large-scale and materials design.

Chemical Reactivity

Partial charges play a crucial role in qualitatively understanding chemical reactivity by identifying sites of electrophilicity and nucleophilicity within molecules. Atoms bearing a partial positive charge (δ+) act as electrophilic centers, attracting s, while those with a partial negative charge (δ−) serve as nucleophilic sites, seeking electrophiles. This guides the approach of reactants in many organic transformations; for instance, in SN2 reactions, the targets the δ+ carbon atom of the substrate, which is polarized by the electronegative , facilitating backside attack and inversion of configuration. Quantitatively, partial charges inform extensions of the , where local partial charges help predict reactivity trends by quantifying the electrostatic contributions to acid-base interactions. In local HSAB models, partial charges on atomic sites correlate with local hardness and softness descriptors, enabling the forecasting of and rate preferences in reactions involving polarized species. These models extend global HSAB principles to site-specific predictions, such as the preference for hard-hard or soft-soft pairings based on charge distributions. A representative example is found in carbonyl compounds, where the partial positive charge on the carbon atom (δ+) enhances its electrophilicity, accelerating reactions such as reductions or Grignard additions. This δ+ charge arises from the difference between carbon and oxygen, polarizing the C=O bond and lowering the activation barrier for nucleophile approach compared to less polarized systems. Furthermore, partial charges correlate with molecular orbital energies to predict in pericyclic reactions, such as Diels-Alder cycloadditions. In these processes, the alignment of δ+ and δ− sites with the coefficients of the highest occupied (HOMO) and lowest unoccupied (LUMO) determines the preferred orientation, as seen in the ortho-para directing effects of substituents on dienes and dienophiles. analysis of partial charges complements orbital , providing electrostatic insights into why electron-withdrawing groups on the dienophile enhance reactivity at the more positive carbon.

Limitations

Accuracy Issues

The assignment of partial charges to carries inherent uncertainties because atomic charges are not directly physical quantities. For molecules, typical uncertainty ranges in partial charge assignments are around ±0.1 e, while for ionic or highly polar compounds, these ranges can extend to ±0.2 e. These uncertainties arise from the choice of population analysis method and the underlying quantum mechanical approximation, limiting the precision of charge-based models in predicting molecular properties. Partial charges are particularly sensitive to the computational level employed, including the choice of basis set and theoretical method. Charges calculated using theory often differ substantially from those obtained with (DFT), with HF typically overestimating charge separation due to lack of electron correlation. Convergence to the complete basis set (CBS) limit is slow for many methods, such as quantum theory of atoms in molecules (QTAIM) and natural population analysis (NPA), where root-mean-square deviations (RMSD) between finite basis sets and the CBS limit can reach 0.012–0.019 e even with augmented quadruple-zeta basis sets. In contrast, methods like Hirshfeld and GAPT converge more rapidly, achieving RMSD below 0.001 e with triple-zeta basis sets. This basis set dependence underscores the need for careful selection of computational parameters to minimize variability in charge values. Validation of partial charge assignments commonly involves comparing derived molecular properties, such as moments, to experimental data or high-level theoretical electron densities. Root-mean-square deviations between partial charges from different methods or levels and reference values are often 0.01–0.10 e, reflecting the inherent ambiguity in charge partitioning. For instance, when benchmarking against experimental moments, charge models at the limit with hybrid DFT functionals yield RMSDs as low as 0.009 e for NPA charges, while pure calculations show higher deviations up to 0.105 e. Post-Hartree-Fock methods, such as second-order Møller-Plesset (MP2), incorporate electron correlation to reduce these errors significantly compared to semi-empirical approaches; MP2 achieves RMSD values 20–80% lower than or semi-empirical methods like AM1 in basis set convergence tests, with typical improvements of 20–30% in reproducing reference charge distributions for diverse molecular systems.

Methodological Challenges

One fundamental methodological challenge in assigning partial charges arises from the inherent ambiguity in partitioning the density of a among its atoms, as there is no unique or physically rigorous way to perform this division. This ambiguity is particularly evident in quantum mechanical population analyses, such as the Mulliken method introduced in the , where charges are derived by equally splitting overlap densities between basis functions centered on different atoms. However, Mulliken charges exhibit strong dependence on the choice of basis set, leading to variations or even sign changes in assigned values upon basis set rotation or expansion, which undermines their reliability for comparative studies. Another significant issue is the limited transferability of partial charges derived from gas-phase calculations to condensed-phase environments, such as solutions, where molecular by surrounding molecules alters the charge distribution. Gas-phase models typically neglect these dynamic effects, resulting in charges that poorly reproduce electrostatic or properties in implicit or explicit simulations. For instance, fixed partial charges optimized in vacuum often overestimate or underestimate moments and energies in aqueous media, necessitating additional corrections like inducible dipoles or rescaling to account for environmental influences. Semi-empirical methods, classified as Class IV charge assignment approaches, further complicate matters by relying on empirical parameterization fitted to experimental properties like moments or electrostatic potentials, which can lead to and reduced generalizability. The Gasteiger-Marsili method, for example, iteratively equalizes electronegativities based on connectivity and empirical parameters to assign charges, but this fitting process risks capturing noise in training data rather than underlying physical trends, especially for diverse molecular classes beyond the parameterization set. This partitioning ambiguity has fueled ongoing debates in since Mulliken's seminal work in the 1950s, with no on a "best" method emerging despite decades of refinement. Recent advances in the 2020s, including models trained on high-level quantum data for data-driven charge assignment, aim to mitigate these issues by learning transferable partitions from large datasets, yet they introduce new concerns regarding interpretability and the "black-box" nature of predictions, potentially obscuring physical insights.

References

  1. [1]
  2. [2]
    Polar Covalent Bonds
    A lowercase Greek delta (δ δ ) is used to indicate that a bonded atom possesses a partial positive charge, indicated by δ+, δ + , or a partial negative charge, ...
  3. [3]
    The Atomic Partial Charges Arboretum: Trying to See the Forest for ...
    Atomic partial charges are among the most commonly used interpretive tools in quantum chemistry. Dozens of different 'population analyses' are in use, ...
  4. [4]
    Quantifying Atomic Volume, Partial Charge, and Electronegativity in ...
    The predictive and explanatory roles of atomic properties such as size, charge, and electronegativity are closely linked to their definitions.
  5. [5]
    Atomic Charges via Electronegativity Equalization - ScienceDirect.com
    The usual way of representing the electronic charge distribution is as partial atomic charges, which are assigned to each atom in a molecule and measured in ...
  6. [6]
    Experimental determination of partial charges with electron diffraction
    Aug 20, 2025 · Abstract. Atomic partial charges, integral to understanding molecular structure, interactions and reactivity, remain an ambiguous concept ...
  7. [7]
    Polarity of Hydrogen Chloride (HCl) - Chemistry Learner
    Therefore, the electronegativity difference results in a partial positive charge on hydrogen and a partial negative charge on chlorine, making HCl a polar ...
  8. [8]
    7.2 Covalent Bonding - Chemistry 2e | OpenStax
    Feb 14, 2019 · The greater the difference in electronegativity, the more polarized the electron distribution and the larger the partial charges of the atoms.
  9. [9]
  10. [10]
    [PDF] Partial charges
    Abstract: Partial charge in theoretical chemistry usually refers to atomic charge, which reflects net charge carried by an atom in a chemical system, ...
  11. [11]
    10.2.2 Atomic Partial Charges - Q-Chem Manual
    By definition, the ChElPG atomic charges are the ones that provide the best fit to the molecular electrostatic potential, evaluated on a real-space grid outside ...
  12. [12]
    Broadband Dielectric Spectroscopy—A Practical Guide
    Feb 24, 2021 · The theory of a molecular dipole was first proposed by P. Debye in 1912 as it applied to charge distributions in molecules ( 1). This work ...
  13. [13]
    Prediction of 13C NMR chemical shifts by artificial neural network. I ...
    Thus, the aim of a current paper is to show which type of partial charges has better correlation with 13C NMR chemical shifts to construct a neural network ...
  14. [14]
    Partial atomic charge of oxygen and hydrogen-bonding ability
    In this work, we present a systematic experimental investigation of hydrogen bonding in the gas phase using mass-selective IR spectroscopy of jet-cooled p- ...
  15. [15]
    Introducing DDEC6 atomic population analysis: part 1. Charge ...
    Apr 13, 2016 · The objective of this article is to develop an atomic population analysis method that is suitable to be used as a default method in quantum chemistry programs.Missing: partial | Show results with:partial
  16. [16]
    Introducing DDEC6 atomic population analysis: part 4. Efficient ...
    Jan 11, 2018 · A key improvement of the DDEC6 method is that it uses a fixed reference ion charge with a total of seven charge partitioning steps to ensure ...
  17. [17]
    10.2.2 Population Analysis: Atomic Partial Charges - Q-Chem Manual
    While Mulliken and Löwdin population analyses are commonly employed, and can be used to produce information about changes in electron density and also localized ...
  18. [18]
    Population | Gaussian.com
    Aug 31, 2022 · This properties keyword controls printing of molecular orbitals and several types of population analysis and atomic charge assignments.
  19. [19]
    5.1. Population Analysis - ORCA 6.1 Manual
    ORCA provides very detailed information about calculated molecular orbitals and bonds through Mulliken, Löwdin, and Mayer population analyses.
  20. [20]
    Development and testing of a general amber force field - Wang - 2004
    Apr 13, 2004 · Here, we describe a general Amber force field that works for most of the pharmaceutical molecules, and which is designed to be as compatible as ...Atom Type Definition · Bond Parameters · Torsional Angle Parameters<|control11|><|separator|>
  21. [21]
    Recent developments and applications of the CHARMM force fields
    Jun 28, 2011 · 35,36 This method allows the values of the partial charges to respond to the electric field of their environment effectively altering the ...
  22. [22]
    Validating CHARMM parameters and exploring charge distribution ...
    Atomic partial charges are the primary components of the electrostatic energy terms and are critical for adequately describing the correct desolvation penalty ...
  23. [23]
    Current Status of the AMOEBA Polarizable Force Field
    We show that the AMOEBA force field is in fact a significant improvement over fixed charge models for small molecule structural and thermodynamic observables ...Introduction · The AMOEBA Force Field · Software Infrastructure for... · Conclusions
  24. [24]
    Current Status of the AMOEBA Polarizable Force Field - PMC
    We show that the AMOEBA force field is in fact a significant improvement over fixed charge models for small molecule structural and thermodynamic observables ...
  25. [25]
    Using Electronegativity to Identify Atoms Bearing Charges after ...
    May 14, 2025 · Most polar organic reactions occur between nucleophiles (atoms or species with a negative ion or partial negative charge) and electrophiles ( ...
  26. [26]
    Steric Effects and Solvent Effects on SN2 Reactions - American ...
    atom partial charges are seen: charge polarizations of 0.03 and. 0.05 au for R1 and R2, respectively, when the solvated density is relaxed at the gas-phase ...
  27. [27]
    Local reactivity descriptors to predict the strength of Lewis acid sites ...
    Lewis acidity of alkali cation-exchanged zeolite is studied using local reactivity descriptors based on hard-soft acid–base (HSAB) concept.
  28. [28]
    Local Softness and Hardness Based Reactivity Descriptors for ...
    The DFT-based reactivity descriptors “local softness” and “local hardness” are used as reactivity indices to predict the reactivity sequences (both ...
  29. [29]
    Carbonyl Group - an overview | ScienceDirect Topics
    We shall see that the carbonyl carbon atom is electrophilic and the carbonyl oxygen atom is nucleophilic.
  30. [30]
    Diels–Alder reactions and electrophilic substitutions with atypical ...
    Nov 6, 2020 · This analysis is based on the frontier molecular orbitals (FMOs), and the natural bond orbital (NBO) analysis (partial charges). Although ...
  31. [31]
  32. [32]
    Population analysis and the effects of Gaussian basis set ... - NIH
    Apr 19, 2023 · In this work, atomic charges were calculated using several population analysis methods including orbital-based methods (Mulliken, Löwdin, and Natural ...Missing: paper | Show results with:paper
  33. [33]
    The SAMPL4 hydration challenge: Evaluation of partial charge sets ...
    IPolQ is motivated by a derivation suggesting that accurate solution-phase energies can be obtained by using partial charges halfway between partial charges ...
  34. [34]
    Explicit Polarization: A Quantum Mechanical Framework for ...
    Aug 6, 2014 · (58) Thus, it appears that condensed phase polarization effects significantly reduce the amount of charge transfer between cations and anions in ...
  35. [35]
    DASH: Dynamic Attention-Based Substructure Hierarchy for Partial ...
    Sep 22, 2023 · The right panel of Figure 7 shows the RMSE values of the DASH partial charges with respect to the MBIS reference charges on the validation set ...