Fact-checked by Grok 2 weeks ago

HSAB theory

HSAB theory, also known as the Hard-Soft Acid-Base theory, is a qualitative framework in coordination chemistry that categorizes as "hard" or "soft" according to their relative , size, and charge density, with the principle stating that hard acids form more stable complexes with hard bases, while soft acids prefer soft bases. Introduced by Ralph G. Pearson in 1963, the theory aims to unify patterns observed in inorganic and by emphasizing electrostatic and covalent interactions in acid-base associations. Hard acids are typically small, highly charged species with low , such as H⁺, Li⁺, and Al³⁺, whereas soft acids are larger, more polarizable entities like Ag⁺, Cu⁺, and Hg²⁺. Similarly, hard bases feature compact, electronegative donor atoms with localized electron pairs, exemplified by F⁻, OH⁻, and NH₃, in contrast to soft bases like I⁻, CN⁻, and PH₃, which have more diffuse electrons. The theory's foundational paper, published in the Journal of the , rapidly gained influence, becoming one of the journal's most cited works and providing a predictive tool for reaction outcomes where traditional acid-base strengths alone were insufficient. Pearson expanded on the concept in subsequent publications, including a 1968 educational review that detailed its applications in estimating acid-base strengths and interpreting equilibrium constants for ligand exchange reactions. In practice, HSAB principles guide the design of stable metal complexes, explain selectivity in solvent extraction processes, and inform mechanisms in , such as the preference of soft Pd²⁺ for soft ligands. While primarily empirical, the theory has been quantified through parameters like absolute hardness (η = (I - A)/2, where I is potential and A is ), linking it to for computational predictions. Despite limitations in cases dominated by steric or , HSAB remains a for understanding reactivity trends across diverse chemical systems.

Historical Development

Origins and Formulation

The conceptualization of HSAB theory emerged from foundational ideas in during the early , which emphasized the sharing of pairs in chemical bonds, and from Irving Langmuir's explorations of acid-base interactions in the 1920s. Langmuir, building on Gilbert N. 's -pair bonding model, described in his 1920 work how atoms or molecules could function as acids by accepting pairs or as bases by donating them, often in the context of achieving stable octet configurations. This donor-acceptor framework provided an early qualitative lens for understanding non-electrostatic interactions in chemical associations, influencing later developments in Lewis acid-base chemistry. By the mid-20th century, scattered observations highlighted inconsistencies in metal-ligand complex stabilities that electrostatic models, such as those based solely on , failed to explain. For instance, certain metal ions showed preferential binding to specific donor atoms—such as halides or pseudohalides—independent of ionic size or charge. A key contribution came in 1958 when Ahrland, Joseph Chatt, and Neville R. classified metal ions into "class a" (preferring hard donors like oxygen or ) and "class b" (favoring soft donors like or iodine), based on empirical trends in coordination compounds. This dichotomy, published in a comprehensive review, underscored the need for a broader theoretical framework beyond traditional acid-base paradigms. The formal proposal of HSAB theory was introduced by Ralph G. Pearson in 1963, who coined the terms "hard" and "soft" to describe exhibiting similar preferences in reactivity. Pearson's seminal paper in the Journal of the presented the hard-soft dichotomy as a qualitative principle to predict the relative strengths of acid-base interactions, motivated by anomalies in stability constants for complexes where ionic models predicted stability but observed preferences favored like-with-like pairings (e.g., hard acids with hard bases). This formulation unified prior classifications, such as Ahrland's, into a cohesive guideline for interpreting acid-base behavior without relying on quantitative alone.

Key Contributors and Evolution

Ralph G. Pearson was the primary architect of HSAB theory, authoring multiple influential papers between 1963 and 1973 that expanded its scope and applications. His foundational 1963 article in the Journal of the formally proposed the of as hard or soft, building on earlier observations of metal ion behaviors to predict stability in acid-base interactions. Pearson further elaborated the core principles in a 1968 publication in the Journal of Chemical Education, emphasizing the qualitative rules governing hard-hard and soft-soft preferences. In 1969, he provided a detailed synthesis of the theory's implications in a review chapter within Survey of Progress in Chemistry, which served as an early compendium for researchers. Robert S. Mulliken's contributions in the 1960s, particularly his development of and concepts, indirectly shaped HSAB by providing a framework for understanding soft acid behaviors through orbital interactions and electron donation-acceptance processes. These ideas, detailed in Mulliken's work on covalent bonding, informed Pearson's extension of HSAB to systems involving polarizable species. Early experimental validations, such as those exploring coordination preferences in metal complexes, reinforced the theory's predictive power during this period. By the late 1960s, HSAB evolved from a predominantly qualitative tool to semi-quantitative methods, incorporating initial empirical scales for acid-base hardness based on experimental stability constants (e.g., log K values) to better quantify preferences. Concurrently, theoretical support emerged from , with Gilles Klopman providing a quantum mechanical interpretation in 1968 that explained HSAB preferences through interactions between the of the base and the of the acid. This progression facilitated broader adoption, with HSAB principles appearing in coordination chemistry textbooks by the early 1970s, such as those emphasizing ligand-metal stability. The 1970s also saw debates on extending HSAB to systems, where its utility in predicting reaction pathways was both championed and critiqued for limitations in covalent contexts.

Core Principles

Classification of Acids and Bases

In the Hard-Soft Acid-Base (HSAB) theory, Lewis acids are classified as hard or soft primarily based on their size, charge, and . Hard acids possess a high charge-to-radius ratio, resulting in high and low polarizability; they are typically small ions in high oxidation states, such as H⁺, Al³⁺, and Fe³⁺. In contrast, soft acids have low and high polarizability, often due to larger size, lower oxidation states, or d¹⁰ electron configurations, exemplified by species like Hg²⁺, Cu⁺, and I⁺. Lewis bases are similarly categorized by the properties of their donor atoms. Hard bases feature donor atoms with high and low , leading to strong electrostatic interactions and a preference for ; representative examples include F⁻, NH₃, and H₂O. Soft bases, however, involve donor atoms that are highly polarizable and of lower , facilitating covalent interactions, as seen in I⁻, CN⁻, and RS⁻ (where R is an ). A comprehensive classification of common acids and bases according to HSAB theory is provided in the following tables, drawn from the foundational work. These lists are not exhaustive but illustrate the key patterns observed in experimental stability constants and reactivity trends.

Hard Acids

Ion/MoleculeExamples
Alkali and alkaline earth metalsH⁺, Li⁺, Na⁺, K⁺, Be²⁺, Mg²⁺, Ca²⁺
Higher oxidation state transition metalsAl³⁺, Sc³⁺, Cr³⁺, Co³⁺, Fe³⁺
Lanthanides and actinidesLa³⁺, Gd³⁺, Th⁴⁺
OthersBF₃, CO₂, Cr(VI), Ti(IV)

Soft Acids

Ion/MoleculeExamples
Low oxidation state coinage metalsCu⁺, Ag⁺, Au⁺, Hg⁺
Other soft metal ionsPd²⁺, Pt²⁺, Cd²⁺, Hg²⁺, Tl³⁺
Halogen and pseudohalogen speciesI⁺, Br⁺, I₂, Br₂, ICN
Organic and boron speciesBH₃, trinitrobenzene, quinones

Hard Bases

Ion/MoleculeExamples
Halides and oxidesF⁻, Cl⁻, O²⁻, OH⁻
Nitrogen and oxygen donors, H₂O, RO⁻ (alkoxides), NO₃⁻
OthersCO₃²⁻, SO₄²⁻, PO₄³⁻

Soft Bases

Ion/MoleculeExamples
Sulfur and heavier chalcogen donorsRS⁻, R₂S, SO₃²⁻, S₂O₃²⁻
Carbon donorsCN⁻, , R₃P, C₆H₅⁻ (phenyl)
OthersI⁻, SCN⁻, R₃As, olefins, H⁻
Many species fall into a borderline category, exhibiting ambiguous or context-dependent behavior that does not strictly align with hard or soft classifications. These include ions such as Zn²⁺, Fe²⁺, Co²⁺, Ni²⁺, , and Pb²⁺ for acids, and bases like , Br⁻, and N₃⁻, which can interact effectively with both hard and soft counterparts depending on conditions. The classification is influenced by several factors beyond intrinsic atomic properties. Hard-hard interactions typically favor due to strong electrostatic attractions, while soft-soft pairs promote more covalent character through orbital overlap and charge transfer. Additionally, effects play a role; in protic solvents like , which act as hard bases, hard acids and bases become effectively harder due to strong solvation shells that stabilize high , whereas soft species are less solvated and retain their .

Fundamental Rules and Predictions

The fundamental principle of HSAB theory, often referred to as the HSAB principle, posits that hard acids form stronger and more stable bonds with hard bases, while soft acids preferentially interact with soft bases. This qualitative rule provides a framework for predicting the affinity between based on their classifications, emphasizing a "like prefers like" tendency in acid-base interactions. Secondary predictions extend this principle to the nature of the resulting bonds and their environmental dependence. Hard-hard interactions typically exhibit greater ionic character due to the high and low of the species involved, rendering them more stable in polar protic solvents like , which can effectively solvate the charged components through hydrogen bonding. In contrast, soft-soft interactions are predominantly covalent, arising from better orbital overlap and electron sharing, and these pairs demonstrate enhanced stability in polar aprotic solvents such as acetone or , where is weaker and does not disrupt the covalent bonding. The underlying stability of matched pairs can be explained through the concept of frontier orbital energy matching. In hard-hard associations, the interaction is largely electrostatic, with minimal orbital overlap, leading to low changes; mismatched pairs, however, incur higher costs due to unfavorable charge transfer or distortion. For soft-soft pairs, the closeness in between the highest occupied (HOMO) of the base and the lowest unoccupied (LUMO) of the acid facilitates efficient covalent bonding without significant charge separation. This avoidance of destabilizing charge transfer in like-with-like pairings underpins the theory's predictive power. Qualitative examples illustrate these rules effectively. Consider the hard acid Na⁺, which forms a more stable interaction with the hard base Cl⁻ than with the soft base I⁻ in aqueous environments due to better of the ionic pair. Ambidentate ligands like (SCN⁻) further demonstrate selectivity: the soft end binds preferentially to soft acids such as Pt²⁺, forming Pt-SCN, while the hard nitrogen end coordinates to hard acids like Co³⁺, yielding Co-NCS.

Quantitative Measures

Chemical Hardness and Softness

In the context of HSAB theory, chemical hardness (\eta) is defined as the resistance of a ' electron cloud to deformation or under the influence of an external or with another . This concept quantifies the stability of the electron distribution, where harder species exhibit lower and greater rigidity in their electronic structure. Conversely, chemical softness (\sigma) is the inverse of hardness, representing the ease with which the electron cloud can be distorted, often associated with higher and greater reactivity toward soft counterparts. Qualitative scales for hardness and softness in HSAB theory are established based on trends in ionization potentials (high for hard species) and electron affinities (low for hard species), reflecting the energy required to alter the . Hard acids and bases typically possess large HOMO-LUMO gaps, indicating low susceptibility to or excitation, whereas soft species feature smaller gaps and higher ease of electron cloud adjustment. Hardness is formally related to electronegativity through operational definitions derived from frontier orbital theory, where absolute electronegativity (\chi) is the average of the ionization energy (I) and electron affinity (E_a), and hardness is half their difference: \eta = \frac{I - E_a}{2} This formulation bridges the qualitative hard-soft dichotomy of HSAB's core principles to a quantitative measure, with I and E_a approximated from vertical ionization processes at constant nuclear configuration. Softness follows directly as \sigma = 1/\eta, emphasizing its role as a measure of global reactivity. While global hardness and softness describe the overall properties of a molecule or ion, local variants apply to specific atomic sites within a molecule, allowing assessment of site-specific interactions in more complex systems.

Calculation Methods and Parameters

Theoretical methods for calculating absolute within the framework of HSAB theory primarily rely on (DFT), where η is approximated from frontier molecular orbital energies as η = (εLUMO - εHOMO)/2, based on . This approximation stems from the exact DFT definition of absolute as η = (1/2)(∂²E/∂N²)v, where E is the total energy, N is the number of electrons, and v is the external potential, but the orbital energies provide a practical computational route for s and s. Early theoretical approaches in the and used semi-empirical methods or Hartree-Fock calculations for εHOMO and εLUMO, but these suffered from limitations such as overestimation of band gaps and poor handling of , leading to less accurate η values before the widespread adoption of DFT in the . Experimental proxies for hardness often derive from ionization potential (IP) and electron affinity (EA), with η ≈ (IP - EA)/2, where IP and EA are measured via electrochemical techniques like to obtain oxidation and reduction potentials, respectively. Spectroscopic methods, such as UV-Vis spectroscopy, provide indirect estimates through excitation energies that approximate the HOMO-LUMO gap (≈ 2η) or assess α, which inversely correlates with since softer species exhibit higher polarizability. Related parameters extend hardness to local and global reactivity descriptors. The Fukui function f(r) = (∂ρ(r)/∂N)v quantifies local reactivity, enabling computation of local softness s(r) = S · f(r), where S = 1/η is the global softness and ρ(r) is the ; this localizes the HSAB principle to specific atomic sites. The electrophilicity index ω = μ²/(2η) further integrates with the μ ≈ (εHOMO + εLUMO)/2, measuring a species' capacity to acquire electrons and aligning with HSAB preferences for hard-hard or soft-soft interactions. Representative computed hardness values illustrate the scale; note that for H⁺, which has no electrons, hardness is theoretically infinite, emphasizing its extreme hardness. The following table summarizes standard absolute hardness values (in eV) for select common species from Pearson's scale, highlighting the hard-to-soft gradient:
SpeciesTypeη (eV)Reference
H⁺
Li⁺35.1
F⁻7.0
Cl⁻4.7
I⁻3.7
CH₃⁺11.0
These values, typically obtained via IP and EA measurements or DFT with basis sets like 6-31G*, underscore quantitative distinctions in HSAB classifications.

Applications and Examples

Reactivity and Stability in Complexes

The HSAB theory provides a for predicting the reactivity and of coordination by emphasizing the preference of hard for hard and soft for soft , leading to higher constants for matching pairs. This is quantitatively observed in log K values, where hard-hard interactions often yield more stable compared to mismatched pairs. For instance, the hard Fe³⁺ forms a highly stable with the hard EDTA, with a log K of approximately 25.1, significantly higher than its with soft ligands such as (log K ≈ 2.2), illustrating the enhanced thermodynamic from electrostatic dominance in hard-hard bonding. The Irving-Williams series, which describes the increasing stability of complexes for divalent first-row transition metals from Mn²⁺ < Fe²⁺ < Co²⁺ < Ni²⁺ < Cu²⁺ > Zn²⁺ with nitrogen or oxygen donors, can be rationalized through HSAB principles. This trend arises from the progressive increase in the softness (or covalent character) of the metal ions from Mn²⁺ to Cu²⁺, enhancing their affinity for borderline bases like amines, before reverting to harder behavior in Zn²⁺; quantitative hardness values refine these predictions by correlating decreasing η (hardness) with rising stability up to Cu²⁺. Solvent effects further modulate complex stability according to HSAB, as protic solvents like act as hard bases that stabilize hard-hard pairs through hydrogen bonding and better solvation of ionic , whereas non-polar media favor soft-soft interactions by reducing electrostatic competition. In aqueous environments, hard acids such as Al³⁺ exhibit enhanced reactivity with hard ligands like , while soft acids like Hg²⁺ show diminished stability; conversely, in aprotic or non-polar solvents, soft-soft pairs like Pd²⁺ with phosphines gain prominence. Practical applications of HSAB in coordination include explaining the of soft metals, where Cd²⁺, a soft acid, preferentially binds to soft sites in biomolecules like (forming Cd-GSH complexes), displacing essential harder metals like Zn²⁺ and causing cellular disruption. Additionally, HSAB guides chromatographic separations of metal ions, where stationary phases with hard oxygen donors selectively retain hard acids (e.g., Ca²⁺), while soft -based phases capture soft acids (e.g., Ag⁺), enabling efficient purification based on differential affinities.

Organic and Inorganic Reactions

In , the HSAB principle guides in nucleophilic substitutions involving ambidentate nucleophiles, where the hard or soft character of the electrophilic site dictates the preferred site of attack. For instance, the thiocyanate ion (SCN⁻), with its hard end and soft end, reacts with primary alkyl halides—considered hard electrophiles—at the nitrogen atom to form alkyl isothiocyanates, while allylic halides, which present softer carbon centers due to stabilization, favor attack at the sulfur atom to yield alkyl thiocyanates. This selectivity arises from the principle that hard-hard and soft-soft interactions are favored in the . Regioselectivity in the alkylation of unsymmetrical s exemplifies HSAB applications, where ambidentate enolate ions (with hard oxygen and softer carbon sites) preferentially alkylate at the less substituted (harder, less sterically hindered) carbon in SN2 reactions with primary or secondary electrophiles, as the resembles a hard-hard interaction. For example, the kinetic enolate of 2-butanone reacts with ethyl bromide primarily at the less substituted methyl carbon, enhancing in . This aligns with HSAB by classifying unhindered alkyl carbons as harder electrophiles that pair better with the enolate's hard oxygen or less polarizable carbon terminus. In inorganic reactions, HSAB predicts outcomes in halide exchange processes, where softer s displace harder ones from soft metal centers. A classic case is the reaction AgCl + I⁻ → + Cl⁻, which proceeds favorably (K ≈ 2 × 10⁶) because iodide, a soft , matches the soft Ag⁺ better than the hard . Similarly, in organomercury systems, CH₃HgCl + I⁻ → CH₃HgI + Cl⁻ has a large (K ≈ 2 × 10³), driven by the soft-soft affinity between and I. The synthetic utility of HSAB is evident in catalyst design for cross-coupling reactions, where soft ligands stabilize soft (0) intermediates, facilitating and steps. For example, (PPh₃), a soft base, coordinates effectively to Pd(0) in Suzuki-Miyaura couplings, enabling efficient aryl-aryl bond formation with turnover numbers exceeding 10⁴ under mild conditions. This matching enhances catalyst longevity and selectivity, as mismatched hard ligands like amines lead to poorer performance with soft Pd centers.

Modifications and Criticisms

Extensions to the Original Theory

In the 1970s, Ho and co-workers extended the HSAB theory through the concept of symbiotic effects, where the presence of mixed hard and soft ligands in a stabilizes borderline acids more effectively than uniform sets of hard or soft s alone. This extension accounts for how the initial binding of one type of alters the effective or softness of the central metal , influencing subsequent ligand affinities. For example, in complexes of borderline metals such as Zn(II) or Cu(II), a combination of hard donors like oxygen and soft donors like or leads to greater overall stability compared to homoleptic arrangements, as the mixed environment balances electrostatic and covalent interactions. The electrostatic-covalent (ECW) model, parameterized by Drago in the 1970s, offers a quantitative refinement to HSAB by decomposing Lewis acid-base bond enthalpies into electrostatic and covalent components via the equation -\Delta H = E_A E_B + C_A C_B where E_A and E_B represent electrostatic parameters for the acid and base, and C_A and C_B capture covalent contributions. Hard acids and bases exhibit dominant E terms due to , while soft pairs emphasize C terms from orbital overlap, enabling predictions of adduct strengths across diverse systems like amine-borane complexes. This parameterization bridges qualitative HSAB classifications with measurable thermodynamic data. Quantum mechanical extensions in the integrated HSAB with (DFT), providing computational tools to assess hardness and softness. Parr's electrophilicity index, \omega = \frac{\mu^2}{2\eta}, where \mu is the electronic and \eta is global hardness, quantifies a ' tendency to accept electrons, aligning soft electrophiles with high \omega values and hard ones with lower values for their preference in reactions. Complementing this, Pearson's maximum hardness principle (MHP) posits that stable molecular configurations maximize hardness under constant chemical potential and external potential, offering a thermodynamic rationale for HSAB matching in ground-state structures, such as in acid-base formations. These DFT-based indices have enabled simulations of reactivity trends without empirical fitting. In the 2000s, HSAB principles were merged with to forecast in pericyclic reactions, particularly through analysis of orbital interactions modulated by local and softness. For instance, in Diels-Alder cycloadditions, the relative softness of the diene's HOMO and dienophile's LUMO dictates ortho or meta orientation, with DFT-derived HSAB reactivity descriptors predicting favored pathways where soft-soft orbital overlaps dominate. This integration has elucidated stereoelectronic control in reactions like [4+2] and [3+2] cycloadditions, enhancing predictive models for synthetic planning.

Limitations and Debates

One significant limitation of HSAB theory lies in the ambiguities arising from borderline species, which do not clearly fit into hard or soft categories and thus exhibit unpredictable reactivity depending on environmental factors such as . For instance, Zn²⁺, classified as a borderline acid, shows varying preferences in different solvents; in aqueous media, enhances its hard character, favoring oxygen donors, while in less polar solvents, it behaves more softly toward ligands, leading to inconsistent complex stability. This solvent-dependent shift highlights how external conditions can override theoretical classifications, complicating predictions for such ions. The theory is often criticized for over-simplification, particularly in handling multi-site molecules or scenarios where effects dominate over enthalpic preferences, and for neglecting detailed overlaps in bonding. In multi-site ligands like ambidentate nucleophiles, HSAB fails to consistently predict site selectivity, as and steric factors intervene beyond hardness matching. Early critiques from the , including those by inorganic chemists like C. K. Jørgensen, emphasized that HSAB overlooks covalent contributions from orbital interactions, such as π-backbonding in complexes, reducing its explanatory power for non-electrostatic interactions. Additionally, when drives association—such as in large, flexible systems—the principle's focus on hardness-softness pairing proves inadequate, as seen in cases where mismatched pairs form due to favorable . Experimental challenges further undermine HSAB's reliability, as quantitative measures of and softness vary significantly across methods, leading to inconsistent classifications and poor reproducibility. values derived from (e.g., via global electrophilicity index) often differ from those based on potentials or experimental constants, creating ambiguity in borderline assignments. Moreover, stable complexes can form from soft-soft mismatches when steric factors dominate, as in systems where oxygen-based hard donors encapsulate soft cations like Ag⁺ through size complementarity and preorganization, stabilizing otherwise unfavorable interactions despite HSAB predictions of instability. These cases illustrate how geometric constraints can enforce , bypassing hardness preferences. In modern debates, HSAB is increasingly viewed as a useful rather than a fundamental principle, with quantum chemical approaches like atoms-in-molecules () theory providing more precise alternatives for analyzing charge transfer and bonding topology. analyses reveal that distributions at critical points better explain reactivity than HSAB's qualitative scale, especially in covalent systems where orbital overlap dominates. Critics argue the theory lacks robust predictive power for , as it primarily addresses thermodynamic stability and often fails to forecast reaction rates in organic substitutions or , where activation barriers depend on geometries rather than ground-state matching. This perspective positions HSAB as an educational tool for qualitative insights but insufficient for quantitative modeling in .

References

  1. [1]
    Hard and Soft Acids and Bases
    ### Summary of HSAB Theory from Abstract and Introduction
  2. [2]
    Hard and soft acids and bases, HSAB, part 1: Fundamental principles
    Using the principles of hard and soft acids and bases to estimate the strength and softness of an acid or base.
  3. [3]
    HARD AND SOFT ACIDS AND BASES | C&EN Global Enterprise
    In that paper, the 13th most cited in the 125-year history of JACS , Pearson proposed a general rule: Hard acids prefer to associate with hard bases, and soft ...Missing: theory original
  4. [4]
    Elucidating the hard/soft acid/base principle - AIP Publishing
    May 18, 2006 · A summary of the acid/base exchange reactions considered in this paper, showing whether the products of the reaction are in accord with the HSAB ...
  5. [5]
    Hard and Soft Acids and Bases - ScienceDirect
    1969, Pages 1-52. Survey of Progress in Chemistry. Hard and Soft Acids and Bases. Author links open overlay panel. RALPH G. PEARSON. Show more. Add to Mendeley.
  6. [6]
    The HSAB Principle — more quantitative aspects - ScienceDirect.com
    The HSAB Principle is reviewed, with special emphasis on common misconceptions of its meaning. The most quantitative way to use it is discussed.
  7. [7]
  8. [8]
    Absolute hardness: companion parameter to absolute electronegativity
    Absolute hardness: companion parameter to absolute electronegativity. Click to copy article linkArticle link copied! Robert G. Parr ...Missing: original | Show results with:original
  9. [9]
    [PDF] Chemical hardness and density functional theory
    For anions, the approximation η = (I – A)/2 was clearly not valid. For molecules there were many values of I available, but fewer values of A. Most common ...
  10. [10]
    Ionization Potential, Electron Affinity, Electronegativity, Hardness ...
    Thus, with χ = (IP + EA)/2 and η = IP − EA, the calculated χ and η are linear combinations of the calculated IP and EA, whereas the “experimental” χ and η are ...
  11. [11]
    Prediction of Electron Energies in Metal Oxides - ACS Publications
    Sep 25, 2013 · the chemical hardness is equivalent to half the value of the band gap, that is, (IP – EA)/2, from the central difference of 1/2(∂2E/∂N2)Z.
  12. [12]
    Local Softness and the Fukui Function | ACS Omega
    Feb 25, 2022 · The Fukui function is designed to characterize how much the electron density changes locally with the global change of the number of electrons.
  13. [13]
    [PDF] Natural Indices for the Chemical Hardness/Softness of Metal Cations ...
    Parr and Pearson used density functional theory (DFT) to calculate absolute hardness of Lewis acids and bases 6. The absolute hardness (η) is the second ...<|control11|><|separator|>
  14. [14]
    Natural Indices for the Chemical Hardness/Softness of Metal Cations ...
    Oct 26, 2017 · For bases B, hardness is identified as the hardness of the species B+. Tables of abs. hardness are given for a no. of free atoms, Lewis acids, ...<|control11|><|separator|>
  15. [15]
    Hard and Soft Acid-Base Behavior in Aqueous Solution: Steric ...
    In this article, the behaviour of HSAB in aqueous solution is addressed. A quantitative scale of hardness-softness for acids and bases is presented also.
  16. [16]
    [PDF] Stability constants of Fe(III) and Cr(III) complexes with dl-2-(2 ...
    The conditional stability constants (logK´ML) of the ML complexes were calculated in terms of pH in the range of 2–12, and compared with the same for EDTA and ...
  17. [17]
    [PDF] Accepted Manuscript - RSC Publishing
    The latter property not only explicates the Irving-Williams series but also rationalizes quantitatively the Pearson's concept of hard and soft acids and bases ...
  18. [18]
    Mysteries of Metals in Metalloenzymes - ACS Publications
    The cell must overcome challenges imposed by the relative stability of complexes formed by divalent transition metals, also known as the Irving–Williams series.
  19. [19]
    Different Behavior of the Histidine Residue toward Cadmium and ...
    Nov 18, 2020 · According to Pearson's theory of hard and soft acids and bases (HSAB theory) ... culture, 0.1 mM cadmium increased the glutathione (GSH) ...
  20. [20]
    [PDF] Separation of transition and heavy metals using stationary phase ...
    HSAB theory,[40] explaining why they are not retained as heavily as ... Chromatographic Separation Of Metal-Ions On A Silica. 421. Column, Anal. Chem ...<|control11|><|separator|>
  21. [21]
    Hard soft acids bases (HSAB) principle and organic chemistry
    Building a high-potential silver–sulfur redox reaction based on the hard–soft acid–base theory. Energy & Fuels 2024, 38 (12) , 11233-11239.
  22. [22]
    Virtual Ligand Strategy in Transition Metal Catalysis Toward Highly ...
    Apr 12, 2023 · This Perspective highlights the virtual ligand strategy, which drastically reduces computational costs for quantum chemical calculations of transition metal ...
  23. [23]
    The ECW Model | Journal of Chemical Education - ACS Publications
    The ECW model is a powerful tool for developing an understanding of donor-acceptor interactions. Most modern inorganic chemistry textbooks discuss the model's ...Missing: URL | Show results with:URL
  24. [24]
    The principle of maximum hardness | Accounts of Chemical Research
    ... Pearson Model. The Journal of Physical ... Initial Hardness Response and Hardness Profiles in the Study of Woodward–Hoffmann Rules for Electrocyclizations.Missing: original | Show results with:original
  25. [25]
    Hard–Soft Interactions in Solvent Extraction with Basic Extractants
    Oct 13, 2021 · This makes cadmium(II) a soft Lewis acid according to the HSAB concept. As a result, cadmium(II) bonds more strongly to softer Lewis bases such ...
  26. [26]
  27. [27]
    Ambident Nucleophilic Substitution: Understanding Non‐HSAB ...
    When orbital relaxation is neglected upon modifying the number of electrons, Equations (6) and (7) indeed reduce to the density of the frontier orbitals.
  28. [28]
    [PDF] Marcus Theory or Hard and Soft Acids and Bases (HSAB) Principle?
    Since the hard cations were found to also react at the soft N-site in many cases, Mayr and coworkers argued against Kornblum's conclusion and further suggested ...Missing: rule | Show results with:rule
  29. [29]
    Pearson's Hard Soft [Lewis] Acid Base Principle - Meta-Synthesis
    Ralph Pearson introduced his Hard Soft [Lewis] Acid Base (HSAB) principle in the early nineteen sixties, and in doing so attempted to unify inorganic and ...
  30. [30]
    The Hard/Soft Acid/Base Rule: A Perspective from Conceptual ...
    Pearson's hard/soft acid/base (HSAB) rule indicates that, all other things being equal, hard acids prefer reacting with hard bases and weak acids with weak ...
  31. [31]
    [PDF] Catalysis Science & Technology - RSC Publishing
    44 Similarly to alkoxides, siloxo ligands can interact with metals with a σ-orbital and two pπ orbitals (Figure. 2). ... C. K. Jorgensen, Inorg. Chem., 1964, 3, ...Missing: criticism | Show results with:criticism
  32. [32]
    HSAB Matching and Mismatching in Selective Catalysis and Synthesis
    Aug 6, 2025 · This 225 reference review summarises the bond energetics, co-ordination modes and reactivity of aldehydes, ketones, imines, enones, amides, ...
  33. [33]
    [PDF] On the Predictive Power of Chemical Concepts - arXiv
    May 16, 2021 · Although its now established roots in quantum chemistry allow for explicit calcu- lations, the HSAB principle should not be mistaken as a ...
  34. [34]
    Predicting the chemical reactivity of organic materials using a ...
    Jul 3, 2020 · This finding suggests that hardness/softness cannot be employed to predict the general reaction kinetics for organic molecules, which, to some ...
  35. [35]
    The hard-soft acid-base (HSAB) principle appraisal - ScienceDirect
    Oct 1, 2020 · In the HSAB principle, soft bases prefer and bind well with soft acids and hard bases prefer and bind well with hard acids [73], [74], [75], [76] ...