Fact-checked by Grok 2 weeks ago

Lewis acids and bases

In 1923, American chemist introduced a groundbreaking theory of acids and bases that expands beyond proton transfer to focus on the donation and acceptance of pairs, providing a more general framework for understanding chemical reactivity. According to Lewis's definition, a Lewis acid is any capable of accepting a pair of electrons, often featuring an electron-deficient center such as a vacant orbital, while a Lewis base is any that can donate a pair of electrons, typically possessing a or . This electron-pair perspective unifies diverse reactions, including those not involving hydrogen ions, and contrasts with the contemporaneous Brønsted-Lowry theory, which limits acids to proton donors and bases to proton acceptors. The Lewis theory's versatility is illustrated through common examples: boron trifluoride (BF₃) serves as a classic Lewis acid due to its empty p-orbital on , readily accepting an electron pair from ammonia (NH₃), a Lewis base with a lone pair on nitrogen, to form the adduct H₃N→BF₃. Similarly, metal cations like Al³⁺ act as Lewis acids by coordinating with molecules or ions as bases, leading to the formation of ions such as [Al(H₂O)₆]³⁺. Neutral molecules like (CO₂) can also function as Lewis acids, reacting with ions (O²⁻) from bases to produce carbonates, demonstrating the theory's applicability to both ionic and molecular systems. The significance of Lewis acid-base theory lies in its broad utility across chemistry, enabling the rationalization of coordination compounds, catalytic processes, and organometallic reactions where traditional acid-base models fall short. It underpins modern concepts like the Hard-Soft Acid-Base (HSAB) principle, which predicts reaction preferences based on the of acids and bases, and plays a crucial role in fields such as inorganic synthesis and . By emphasizing electron-pair interactions without requiring involvement, the theory facilitates the design of Lewis acid catalysts, such as AlCl₃ in Friedel-Crafts alkylations, enhancing efficiency in organic transformations.

Fundamentals

Definition and core principles

In 1923, Gilbert N. Lewis proposed a foundational theory of acid-base chemistry centered on electron-pair interactions, defining a Lewis base as a substance that donates an electron pair and a Lewis acid as a substance that accepts an electron pair. This definition shifts the focus from proton transfer to the formation of coordinate covalent bonds, where the shared electron pair originates entirely from the base. The core principle of Lewis acid-base theory lies in the electron-transfer mechanism that results in adduct formation through dative bonds, also known as coordinate bonds, in which both electrons in the bond are provided by the Lewis base. These interactions emphasize the role of vacant orbitals in the acid and lone pairs or pi electrons in the base, enabling a broader scope of reactions beyond those involving hydrogen ions. The general reaction can be represented as: \text{A (acid)} + \text{:B (base)} \to \text{A} \leftarrow \text{B (adduct)} where the arrow indicates the donation of the electron pair from the base to the acid. This electron-pair framework provides applicability to diverse chemical systems, including non-aqueous solvents and reactions without proton involvement, such as the classic example of boron trifluoride (BF_3) acting as a Lewis acid by accepting an electron pair from ammonia (NH_3) to form the adduct H_3N\rightarrowBF_3.

Adduct formation and depiction

In Lewis acid-base chemistry, adduct formation occurs when a Lewis base donates an electron pair from its highest occupied molecular orbital () to the lowest unoccupied molecular orbital (LUMO) of a Lewis acid, resulting in the creation of a , also known as a dative bond. This process involves the base acting as a and the acid as an , leading to a new molecular entity where the shared originates entirely from the base. The dative bond is distinct from traditional covalent bonds in its formation mechanism but is otherwise similar once formed. Common depictions of Lewis acid-base adducts emphasize the directionality of electron donation. The arrow notation, where an arrow points from the donor atom of the base to the acceptor atom of the acid (e.g., B → A), is widely used to illustrate the dative bond and highlight the unidirectionality of electron flow. Alternatively, dashed lines represent the coordinate bond to indicate its partial ionic character, while in some structural contexts—particularly after or hybridization considerations—a full covalent line may be employed to denote the equilibrated bond. A representative example is the adduct between and aluminum trichloride, depicted as \ce{Me3N -> AlCl3}, where the lone pair donates to the empty p-orbital on aluminum, forming a often written with the arrow to show the dative . The stability of these adducts is influenced by several key factors. Steric hindrance from bulky substituents on either the acid or base can weaken the interaction by increasing the distance between the donor and acceptor sites, reducing orbital overlap. Electronic effects, such as the polarity of the acid's LUMO and the basicity of the base's , determine the strength of the donation, with better energy matching leading to stronger bonds. Solvent effects further modulate stability; coordinating solvents can compete with the base for the acid's coordination sites, forming solvates that destabilize the , whereas non-coordinating solvents enhance association. Modern computational approaches provide deeper insights into adduct formation through molecular orbital diagrams that visualize electron density redistribution. These diagrams illustrate the HOMO-LUMO interaction as a stabilizing overlap, where the base's filled orbital donates density into the acid's empty orbital, often quantified via (DFT) calculations to predict bond energies and geometries. Such representations reveal nuances like charge transfer and polarization effects that traditional notations overlook, aiding in the design of novel .

Lewis Acids

Characteristics and simple examples

Lewis acids are defined as electron-pair acceptors that form coordinate covalent bonds with bases by utilizing vacant orbitals or incomplete electron octets on their central atoms. These species are typically electrophilic, featuring central atoms—often from elements like or aluminum—that are bonded to electronegative ligands, resulting in and the availability of low-energy empty orbitals for electron acceptance. This distinguishes Lewis acids from Brønsted acids, as the former do not necessarily involve proton transfer but rather dative bond formation. Simple examples of Lewis acids include neutral molecular compounds with central atoms possessing fewer than eight valence electrons. (BF₃) exemplifies this, where the atom has six valence electrons arranged in three σ-bonds to atoms, leaving an empty 2p orbital perpendicular to the molecular plane for accepting an . Similarly, aluminum trichloride (AlCl₃) features a central aluminum atom with six valence electrons, enabling it to act as an electron-pair acceptor despite its larger atomic size compared to . (BCl₃) shares analogous properties, with ligands providing less effective π-backbonding than , enhancing its acidity relative to BF₃. These simple Lewis acids readily form adducts with electron-pair donors, demonstrating their reactivity. A representative reaction is the combination of BF₃ with (NH₃), where the nitrogen lone pair coordinates to the empty boron orbital: \ce{BF3 + :NH3 -> H3N -> BF3} This dative bond formation stabilizes the electron-deficient boron center, and the adduct's stability depends on the acid's ability to accept the pair without significant steric repulsion. Reactivity trends among simple Lewis acids highlight the influence of substituents on and orbital availability. For boron trihalides, Lewis acidity increases in the order BF₃ < BCl₃ < BBr₃ < BI₃, counter to the electronegativity trend of the halogens; this reversal arises because fluorine's small size and high electronegativity enable strong π-backbonding into the empty boron p-orbital, partially filling it and reducing BF₃'s acidity compared to the heavier halides where such donation is weaker. In contrast, for group 13 trihalides, larger central atoms like aluminum generally confer greater acidity than smaller ones like due to weaker metal-halogen bonds and reduced back-bonding, as seen in AlCl₃ being a stronger acid than BF₃ toward many bases. Steric bulk from ligands can further modulate reactivity, with more crowded acids forming weaker or less stable adducts due to hindered base approach.

Complex and transition metal examples

Transition metal complexes serve as prominent Lewis acids due to their central metal ions possessing partially filled d-orbitals, which enable acceptance of electron pairs from Lewis bases, often leading to high coordination numbers and stable adduct formation. These characteristics allow transition metals to exhibit variable oxidation states and geometries, facilitating diverse coordination chemistries beyond simple main-group acids. Representative examples include trivalent iron (Fe³⁺) and divalent copper (Cu²⁺) ions, which act as Lewis acids by coordinating with ligands such as water or halides to form aquo or chloro complexes, respectively, due to their electron-deficient d-orbitals. Similarly, the hexaamminecobalt(III) complex, [Co(NH₃)₆]³⁺, can demonstrate Lewis acidity at the metal center through interactions in substitution reactions, where additional ligands can approach the coordination sphere. A distinctive feature in such systems is back-bonding, as observed in metal carbonyls like (Ni(CO)₄), where the transition metal acts as a Lewis acid by accepting σ-electron pairs from the carbon monoxide (CO) ligands, while the metal d-orbitals donate electrons back to the π* orbitals of CO, strengthening the overall metal-ligand interaction. Coordination sphere expansion exemplifies this acidity, as in the reaction of the square planar [PtCl₄]²⁻ complex with chloride to form the octahedral [PtCl₆]²⁻, illustrating how the Pt(II) center accepts additional electron pairs to achieve higher coordination. Recent advancements in the 2020s have incorporated transition metals into (FLPs), where sterically hindered metal centers, such as those in iron or nickel complexes, function as Lewis acids alongside bases to activate small molecules like H₂ or CO₂ without traditional adduct formation, enhancing catalytic efficiency in hydrogenation and reduction processes.

Role of H+ and other cations

The proton (H⁺) exemplifies the quintessential Lewis acid, possessing no valence electrons and thus an empty valence shell capable of accepting an electron pair from a to form a coordinate bond. This electron deficiency makes H⁺ the "perfect" Lewis acid, as it readily forms adducts by coordinating to lone pairs on donor atoms. A classic example is its reaction with water, where the proton accepts an electron pair from the oxygen lone pair to generate the : \text{H}^{+} + \text{H}_2\text{O} \rightarrow \text{H}_3\text{O}^{+} Similarly, H⁺ coordinates to the nitrogen lone pair in ammonia, yielding the ammonium ion: \text{H}^{+} + \text{NH}_3 \rightarrow \text{H}_3\text{N}-\text{H}^{+} Although H⁺ represents ultimate electron deficiency, it rarely exists in isolation and is typically solvent-coordinated, as in the hydronium species, which itself acts as a Lewis acid in further interactions. Other cations, such as those from alkali and alkaline earth metals, also function as Lewis acids by accepting electron pairs into vacant orbitals or via electrostatic attraction enhanced by their charge density. Notable examples include Li⁺ in organolithium reagents (RLi) and Mg²⁺ in Grignard reagents (RMgX), where the metal cations coordinate to Lewis bases like ether solvents or functional groups on substrates, stabilizing the reagent and directing reactivity. In these organometallic contexts, the cations' Lewis acidity facilitates adduct formation, such as Mg²⁺ binding to carbonyl oxygens during nucleophilic additions. The strength of this Lewis acidity for such cations correlates directly with their hydration energies, which quantify the exothermic interaction with water lone pairs and increase for smaller ions with higher charges due to greater charge density. This relationship highlights how solvation tendencies parallel the cations' ability to form Lewis acid-base adducts in non-aqueous environments.

Lewis Bases

Characteristics and simple examples

Lewis bases are defined as electron-pair donors that form coordinate covalent bonds with Lewis acids by providing lone pairs or π electrons, typically from nucleophilic atoms such as nitrogen, oxygen, or halogens. These species possess available electron pairs in lone pairs or π bonds, enabling them to act as nucleophiles in dative bond formation. This electron donation distinguishes Lewis bases from , which are limited to proton acceptors, allowing the Lewis concept to encompass a broader range of reactions. Simple examples of Lewis bases include neutral molecules featuring lone pairs on central atoms. Ammonia (NH₃) is a classic Lewis base, with its nitrogen lone pair available for donation to electron-deficient centers. Water (H₂O) similarly functions as a Lewis base via oxygen lone pairs, often coordinating to metal cations. Anionic species like the hydroxide ion (OH⁻) serve as strong Lewis bases, donating electron pairs from oxygen to form hydroxo complexes. These simple Lewis bases readily form adducts with electron-pair acceptors, illustrating their reactivity. A representative reaction is the coordination of ammonia to boron trifluoride (BF₃): \ce{BF3 + :NH3 -> H3N -> BF3} In this adduct, the dative bond from to stabilizes the electron-deficient center, with adduct stability influenced by the base's electron-donating ability and minimal steric hindrance. Reactivity trends among simple Lewis bases are governed by factors affecting electron-pair availability, such as inductive effects and . For amines, Lewis basicity generally increases with alkyl substitution (NH₃ < primary < secondary < tertiary) due to electron-donating inductive effects that enhance lone-pair density on nitrogen. Oxygen-based bases like alcohols exhibit slightly higher basicity than ethers owing to reduced stabilization in the latter, though both are weaker donors than amines. Steric bulk around the donor atom can diminish reactivity by impeding approach to the acid, leading to less stable adducts with bulky Lewis acids.

Complex and ambidentate examples

Complex Lewis bases encompass multidentate ligands that can coordinate to a metal center through multiple donor atoms, forming stable chelate rings, as well as ambidentate ligands that exhibit regioselective binding via different donor sites on the same molecule. Multidentate coordination enhances complex stability through the , where the entropy gain from releasing fewer solvent molecules outweighs the entropic cost of ligand preorganization. Ambidentate behavior arises from the electronic asymmetry of the ligand, allowing donation from alternative atoms depending on the metal's hardness, charge, or steric environment. A classic example of a bidentate Lewis base is ethylenediamine (en, H₂NCH₂CH₂NH₂), which donates electron pairs from both nitrogen atoms to form five-membered chelate rings with transition metals. This ligand's dual donation capability is exemplified in the formation of tris(ethylenediamine)nickel(II), where three en molecules coordinate to Ni²⁺, resulting in an octahedral complex with enhanced thermodynamic stability compared to monodentate analogs. $3 \ce{en} + \ce{Ni^{2+}} \rightarrow [\ce{Ni(en)3}]^{2+} Cyanide (CN⁻) serves as a prototypical ambidentate Lewis base, capable of binding through either the carbon atom (as cyanides, M–C≡N) or the nitrogen atom (as isocyanides, M–N≡C), with the preference influenced by the metal's π-acceptor ability and the complex's overall electronics. Carbon-bound cyanides are common in low-valent, soft metal centers like Fe or Co, facilitating strong σ-donation and π-backbonding, while nitrogen-bound forms appear in harder, higher-oxidation-state metals. Beyond σ-donor lone pairs, certain unsaturated hydrocarbons exhibit π-basicity as Lewis bases, donating π-electron density from alkenes or aromatic rings to electron-deficient metals via the Dewar–Chatt–Duncanson (DCD) model. In this framework, the olefin acts as a σ-donor through its π-orbital while accepting back-donation into its π* orbital, stabilizing organometallic complexes like Zeise's salt (K[PtCl₃(η²-C₂H₄)]). Aromatic systems, such as benzene in (η⁶-C₆H₆)Cr(CO)₃, follow a similar synergistic bonding pattern, where the arene's delocalized π-cloud behaves as a six-electron donor. Post-2010 developments in supramolecular chemistry have introduced advanced ambidentate ligands that enable dynamic, switchable coordination for responsive materials and sensors. For instance, hybrid N-heterocyclic carbene–β-diketonate (IMes-acac) ligands fuse a strong σ-donor carbene with an ambidentate O,O'-chelate, allowing selective metallation at N or O sites to form heteropolymetallic assemblies with tunable redox properties. These ligands exploit ambidenticity for self-assembly into cages or networks, as seen in luminescent first-row transition metal complexes where ligand flipping modulates emission wavelengths. Such innovations extend Lewis basicity concepts to programmable architectures beyond traditional coordination spheres.

Role of anions and lone-pair donors

Anions function as Lewis bases primarily through the donation of lone pairs of electrons to electron-deficient Lewis acids, resulting in the formation of ionic adducts that enhance the stability of the resulting species. Halide ions, such as chloride (Cl⁻), exemplify this role by coordinating to metal centers with incomplete octets; for instance, the reaction of chloride with aluminum chloride yields the tetrahedral tetrachloroaluminate anion, where Cl⁻ donates its lone pair to the empty orbital on Al: \ce{Cl^- + AlCl3 -> AlCl4^-} This interaction underscores the utility of halide anions in expanding coordination spheres and stabilizing reactive intermediates. Similarly, the oxide ion (O²⁻) serves as a potent electron-pair donor, as demonstrated in the formation of carbonates from metal oxides and carbon dioxide, where O²⁻ binds to the electrophilic carbon atom of CO₂ to produce stable ionic compounds like CaCO₃. The hydride ion (H⁻) also acts as an anionic Lewis base, leveraging its lone pair to interact with acids like boranes or metal cations, often leading to hydride transfer or complex formation due to its strong reducing character. In many anions, the lone-pair donation originates from heteroatoms bearing the negative charge, amplifying the basicity compared to their neutral counterparts. The hydroxide ion (OH⁻), for example, donates a from its oxygen atom to Lewis acids such as aluminum or compounds, forming hydroxo complexes that are integral to aqueous coordination ; the oxygen in OH⁻ possesses three lone pairs, enabling versatile binding modes. This donation mechanism parallels that in other oxyanions but is particularly pronounced in simple cases like OH⁻, where the charge concentrates on oxygen for efficient overlap with acid orbitals. Certain anions exhibit extraordinary basicity, classifying them as superbases capable of deprotonating even weak acids. The amide ion (NH₂⁻), with its nitrogen-centered , exemplifies this, displaying basicity far exceeding that of due to the lower of and resulting higher electron availability; its pK_b value approaches -35 in some solvents, enabling reactions like the of terminal alkynes. Unlike adducts from neutral bases, those involving anions like halides, oxides, hydrides, hydroxides, and amides typically manifest as stable salts or ionic lattices rather than discrete molecular species, owing to the electrostatic reinforcement of the coordinate bonds in charged environments.

Classification Systems

Hard-soft acid-base (HSAB) theory

The hard-soft acid-base (, proposed by G. Pearson in 1963, classifies Lewis acids and bases into "hard" and "soft" categories based on their and , predicting that hard acids preferentially interact with hard bases, while soft acids prefer soft bases, leading to more stable . Hard species are characterized by low , high , and small size, favoring ionic interactions, whereas soft species exhibit high , lower , and a tendency toward covalent bonding. This qualitative principle extends the Lewis acid-base framework by providing a for understanding reactivity patterns without relying on proton transfer, emphasizing electronic complementarity in formation. Typical hard Lewis acids include highly charged, small cations such as H⁺, Al³⁺, and Fe³⁺, as well as molecules like BF₃ and SO₃, which have empty orbitals with low . Hard bases feature localized lone pairs and high , exemplified by F⁻, OH⁻, NH₃, and H₂O. In contrast, soft acids are larger with diffuse orbitals, such as Cu⁺, Ag⁺, Hg²⁺, and Pd²⁺, while soft bases include highly polarizable donors like I⁻, CN⁻, RS⁻, and phosphines (PR₃). These classifications are derived from empirical , where hard-hard pairings, such as the strong binding in [AlF₆]³⁻, yield thermodynamically stable ionic products, and soft-soft pairings, like [HgI₄]²⁻, form covalent, kinetically inert species. The HSAB principle's predictive power lies in forecasting reaction outcomes and complex stability; for instance, a soft acid like Ag⁺ forms a highly stable complex with the soft base CN⁻ in [Ag(CN)₂]⁻, whereas mismatched hard-soft interactions, such as H⁺ with I⁻, result in weaker, more reactive bonds prone to displacement. A related concept, chemical , explains deviations where the effective of a site is influenced by surrounding ligands; for example, the soft Pd²⁺ in trans-[Pd(NH₃)₂Cl₂] prefers the borderline Cl⁻ over I⁻ due to the hard NH₃ ligands enhancing the site's hardness. This principle guides synthetic strategies in coordination chemistry, such as selective ligand exchange in , where soft metal centers facilitate reactions with soft substrates. Recent computational extensions have quantified HSAB concepts using (DFT), defining chemical η as η = (I - A)/2, where I is ionization potential and A is , to validate qualitative predictions through orbital overlap analyses and energy decompositions. For instance, DFT studies on complexes in the 2020s confirm that hard-hard interactions maximize electrostatic contributions, while soft-soft favor charge-transfer stabilization, with applications in modeling active sites and materials design. These approaches, building on Parr and Pearson's hardness framework, demonstrate HSAB's enduring utility in interpreting electronic effects across diverse systems.

Other classification schemes

In addition to qualitative frameworks like , several quantitative and alternative classification schemes have been developed to describe Lewis acid-base interactions based on thermodynamic, transfer-based, and computational criteria. These approaches provide tools for predicting strengths, behaviors, and reactivity patterns by emphasizing specific mechanisms. One prominent thermodynamic classification is the Drago-Wayland model, which quantifies the of Lewis acid-base formation by separating contributions from electrostatic and covalent ing. In this scheme, each Lewis acid is characterized by parameters E_A (electrostatic) and C_A (covalent), while each base has corresponding E_B and C_B values; the negative change for formation is given by the equation -\Delta H = E_A E_B + C_A C_B where the first term represents ionic or electrostatic interactions and the second covalent or orbital overlap contributions. This double-scale equation allows for the prediction of energies in gas-phase or non-aqueous environments and has been applied to classify acids and bases based on their relative reliance on these bonding types, such as classifying BF_3 as predominantly electrostatic (high E_A, low C_A) compared to more covalent acids like I_2. The Lux-Flood theory offers a specialized classification centered on oxide ion (O^{2-}) transfer, defining a Lewis acid as an oxide acceptor and a base as an oxide donor, particularly useful for high-temperature or molten salt systems involving oxides and oxyanions. For example, SiO_2 acts as an acid by accepting O^{2-} from a base like CaO to form CaSiO_3, enabling classification of metal oxides as basic (oxide donors) and non-metal oxides as acidic (oxide acceptors) in solid-state chemistry. This scheme complements the general Lewis definition by focusing on anionic transfer reactions in contexts where proton or electron transfer is irrelevant. For solvent classification, Gutmann's donor-acceptor number (DN/AN) system measures Lewis basicity and acidity through empirical scales derived from spectroscopic and calorimetric data. The donor number (DN) quantifies a solvent's ability to donate pairs to a reference Lewis acid like SbCl_5, expressed as the negative of the interaction in kcal/mol (e.g., DN = 29.8 for , indicating basicity), while the acceptor number (AN) assesses electrophilicity via ^{31}P NMR shifts of triethylphosphine oxide in the solvent (e.g., AN = 19.2 for ). These parameters classify solvents by their coordination tendencies toward cations (high DN) or anions (high AN), aiding in the selection of media for Lewis acid-base reactions. Recent computational schemes leverage (DFT) to classify Lewis acids and bases using orbital hardness descriptors, extending traditional concepts to molecular-level predictions. In these approaches, global and local (\eta) values, derived from the second of with respect to number, quantify resistance to deformation; for instance, high orbital hardness correlates with "hard" Lewis acids like Al^{3+}, while low values indicate softer ones like Cu^{+}. Post-2015 developments have integrated local softness functions to map site-specific reactivity in complexes, enabling classification of ambidentate ligands or frustrated Lewis pairs based on frontier orbital interactions.

Quantification Methods

Measuring Lewis acidity

The Gutmann acceptor number (AN) provides an experimental scale for quantifying acidity through coordination-induced shifts in the ^{31}P NMR spectrum of triethylphosphine oxide (Et_3PO), a reference Lewis base. In the Gutmann-Beckett method, the chemical shift difference (Δδ) between the free Et_3PO (δ ≈ 41 ppm in CDCl_3) and the acid-base adduct is used to calculate AN = 2.21 × Δδ, normalized such that AN = 0 for non-coordinating solvents like hexane and AN = 100 for SbCl_5 in dilute 1,2-dichloroethane solution. This scale reflects the ability of the Lewis acid to withdraw electron density from the phosphorus , with higher AN values indicating stronger acidity. For example, BF_3 exhibits an AN of 89, while SbCl_5 serves as the reference with AN = 100; other halides like AlCl_3 have AN = 87, demonstrating the method's sensitivity to electronic and in group 13 and 15 Lewis acids. Another experimental approach involves the fluoride ion affinity (FIA), a gas-phase thermochemical measure of Lewis acidity defined as the negative enthalpy change (-ΔH) for the reaction LA + F^- → LA-F^-, where LA is the Lewis acid. For neutral Lewis acids of the form MX_n (M = central atom, X = ligand), FIA can be assessed via the fluoride exchange reaction MX_n + F^- → MF_n + X^-, providing insight into the relative stability of fluorinated adducts and thus the acid's affinity for hard nucleophiles like fluoride. This metric is particularly useful for comparing superacids, with higher FIA values (e.g., 493 kJ/mol for SbF_5) indicating stronger Lewis acids capable of stabilizing anionic adducts. FIA values are often computed or measured for bare cations or neutral species to rank acidity independent of solvent effects, emphasizing hard Lewis acid behavior. Computational methods offer base-independent quantification of Lewis acidity, such as the electrophilicity (ω) derived from (DFT). Introduced by Parr and coworkers, ω measures the stabilization energy of a system upon gaining from a donor, calculated as \omega = \frac{\mu^2}{2\eta} where μ is the electronic (μ = (ε_H + ε_L)/2) and η is the (η = ε_L - ε_H), with ε_H and ε_L being the energies, respectively. Higher ω values correlate with greater Lewis acidity; for instance, BF_3 has ω ≈ 3.5 eV, while stronger acids like B(C_6F_5)_3 exceed 4 eV, allowing prediction of reactivity trends across diverse Lewis acids without experimental probes. This integrates global reactivity descriptors and has been applied to main-group and species for conceptual understanding of acceptance. Kinetic scales provide complementary measures of Lewis acidity through reaction rates with reference nucleophiles. Mayr's electrophilicity parameters () quantify the reactivity of Lewis acids (often as electrophiles) toward a family of benzhydrylium ions or carbocations, where log k = s(N + ) relates second-order rate constants (k) to the nucleophilicity (N) and electrophilicity () of the partners, with s as a nucleophile-specific . Higher values denote stronger Lewis acids; for example, Michael acceptors or iminium ions with > 0 exhibit pronounced acidity toward π-nucleophiles, offering an inverse perspective via nucleophilicity parameters to assess acid strength in transformations. This approach emphasizes dynamic aspects of acidity beyond measures.

Measuring Lewis basicity

Lewis basicity, which quantifies the of a species to donate an , is measured through various experimental and computational s that assess affinity toward reference Lewis acids or probes of . One prominent calorimetric scale is the Gutmann donor number (DN), defined as the negative change (in kcal/) for the formation of a 1:1 between the Lewis base and (SbCl₅) in solvent. This isolates the donor by using the highly electrophilic SbCl₅ as a reference acid and to minimize solvent interference. For instance, exhibits a DN of 18 kcal/, reflecting its moderate basicity, while shows a stronger DN of 59 kcal/ due to the higher availability of its . Hydrogen bond basicity scales provide another experimental approach, particularly for assessing interactions where the base acts as a hydrogen bond acceptor, akin to Lewis donation in proton-coupled processes. The pK_{BHX} scale, for example, measures the negative logarithm of the equilibrium constant for hydrogen bonding between the base and a reference donor like 4-fluorophenol in an inert solvent such as CCl₄, yielding values that correlate with overall Lewis basicity for oxygen and nitrogen donors. This scale spans a range of about 9-10 units, with ethers around pK_{BHX} ≈ 2-3 and amines up to 4, emphasizing steric and electronic factors in lone pair accessibility. Gas-phase measurements offer insight into intrinsic Lewis basicity, free from solvation effects, with proton affinity (PA) serving as a key metric for the tendency to bind H⁺ as a prototypical acid. The PA of a base B is defined as the negative change for the reaction: \mathrm{B(g) + H^{+}(g) \rightarrow BH^{+}(g)} \quad \Delta H = -\mathrm{PA(B)} Values are typically reported in kJ/mol and range from 600-2000 kJ/mol, with at 854 kJ/mol illustrating strong intrinsic basicity compared to at 691 kJ/mol. These affinities are determined via or equilibrium techniques and relate directly to Lewis basicity since H⁺ accepts the electron pair without complicating covalent bonds. Computationally, (DFT) descriptors like local softness (s) quantify site-specific basicity by evaluating responsiveness at potential donor atoms. Local softness at site k for nucleophilic attack (relevant to basicity) is given by s_k^- = S \cdot f_k^-, where S is the global softness and f_k^- is the condensed Fukui function, approximated as the electron density difference between neutral and anionic states. High s_k^- values indicate soft, basic sites prone to Lewis acid coordination, as seen in amines where s_k^- exceeds that of carbonyl oxygens. This approach, rooted in conceptual DFT, aids in predicting reactivity without experiments. Recent advances in have enabled probes of accessibility, a key aspect of Lewis basicity, by monitoring vibrational shifts upon formation. In attenuated total reflection IR, for instance, detects changes in carbonyl or stretching frequencies when bases coordinate to reference acids, revealing steric hindrance or electronic delocalization effects on lone pair availability. Studies from the 2020s, such as those on frustrated Lewis pairs, use these shifts (e.g., >50 cm⁻¹ for N-H or C=O modes) to quantify effective basicity in , complementing traditional scales with dynamic structural insights.

Applications

In catalysis and synthesis

Lewis acids play a pivotal role in by coordinating to substrates, thereby activating electrophilic centers and facilitating nucleophilic attack. In the Friedel-Crafts alkylation, aluminum chloride (AlCl₃) serves as a classic Lewis acid catalyst, coordinating to the carbonyl oxygen of an to generate a highly electrophilic acylium that undergoes . This coordination enhances the electrophilicity of the substrate, enabling efficient C-C bond formation under mild conditions. Lewis bases, conversely, activate nucleophilic species in catalytic cycles. Tertiary amines, such as 1,4-diazabicyclo[2.2.2]octane (), act as Lewis bases in the Baylis-Hillman reaction, where they form zwitterionic intermediates with activated s like acrylates, promoting addition to aldehydes to yield α-methylene-β-hydroxy carbonyl compounds. This mechanism underscores the role of lone-pair donation in generating reactive enolates for stereoselective synthesis. Frustrated Lewis pairs (FLPs), combinations of sterically hindered Lewis acids and bases that avoid adduct formation, have revolutionized metal-free , particularly for small molecule activation. Pioneered by Douglas Stephan in 2006, FLPs comprising bulky phosphines and boranes heterolytically cleave dihydrogen (H₂) at ambient conditions, enabling of imines and other unsaturated substrates without transition metals. This highly cited discovery has influenced developments in sustainable , including CO₂ reduction and alkene hydrofunctionalization. Bifunctional catalysts integrating Lewis acid and base sites offer synergistic activation of dual reaction partners, enhancing selectivity in transformations. These systems, often featuring cooperative metal centers and s, promote tandem reactions like aldol-Michael additions by simultaneously polarizing electrophiles and nucleophiles. Such designs mimic enzymatic efficiency, achieving high yields in asymmetric syntheses. Recent advances in the 2020s have focused on chiral Lewis acids for enantioselective , improving stereocontrol in complex assembly. For instance, chiral Lewis acids have been used to catalyze asymmetric Friedel-Crafts alkylations of indoles, delivering products with high enantiomeric excess and broad substrate scope. These developments, building on ligand optimization, address challenges in pharmaceutical synthesis by enabling scalable, green processes. A representative activation mechanism in Lewis acid catalysis is depicted below, where AlCl₃ coordinates to an aldehyde: \mathrm{RCHO + AlCl_3 \rightarrow RCH=O \leftarrow AlCl_3} This complex lowers the energy barrier for , as seen in aldol reactions.

In coordination chemistry and materials

In coordination chemistry, acid-base interactions underpin the formation of coordination compounds, where metal ions or other electron-pair acceptors serve as acids coordinating to ligands acting as bases, resulting in stable structures such as [M(L)_n] complexes. These interactions enable the assembly of coordination polymers, in which multidentate ligands bridge metal centers to form extended frameworks with tunable and functionality. For instance, Zn^{2+} ions, functioning as acids, coordinate to groups on linkers like terephthalate, yielding metal-organic frameworks (MOFs) such as , which exhibit high surface areas exceeding 3000 m^2/g for gas storage and separation applications. Zeolites, materials with crystalline microporous structures, act as solid Lewis acids through extra-framework aluminum species or exchanged metal cations, facilitating selective processes. These Lewis acid sites interact with incoming cations or polar molecules, enabling applications in and nuclear waste remediation by preferentially binding heavy metal ions like Cs^+ or Sr^{2+}. The capacity of zeolites, often reaching 2-5 meq/g, stems from the framework's negative charge balanced by exchangeable cations, with Lewis acidity enhancing selectivity under mild conditions. In covalent organic frameworks (COFs), Lewis base functionalities incorporated into organic linkers, such as or groups, play a crucial role in framework stability and post-synthetic modification. These nitrogen-containing linkers coordinate to metal ions or hydrogen-bond donors, enabling the design of COFs for selective adsorption; for example, the DhaTph COF uses triazine-based Lewis base sites to capture Th^{4+} ions with distribution coefficients over 10^4 mL/g. Such interactions expand COFs' utility in membranes and sensors, distinct from their metal-free covalent bonding. Recent advancements leverage acid salts in electrolytes to enhance performance in solid-state devices. In 2025, multivalent metal cations like Al^{3+} or Zn^{2+} as Lewis acidic additives promote in-situ gelation of ether-based solvents, forming stable with ionic conductivities above 10^{-3} S/cm at and enabling metal batteries to cycle over 500 times at 4 V. These salts coordinate to anion receptors, suppressing growth and improving oxidative stability up to 4.8 V, addressing key challenges in high-energy-density storage.

Historical Development

Origins and Lewis's formulation

In 1923, Gilbert N. Lewis formulated his theory of acids and bases in the book Valence and the Structure of Atoms and Molecules, proposing that acid-base reactions involve the sharing of an between an acid and a base. In this electronic theory of valence, Lewis described acids as substances capable of employing an from another molecule to complete the stable octet of one of their own atoms, while bases provide that . This definition shifted the focus from ionic dissociation or proton transfer to the fundamental process of electron-pair donation and acceptance in chemical bonding. Lewis's motivation arose from the constraints of the contemporaneous Arrhenius and Brønsted-Lowry theories, which emphasized proton production or transfer and struggled to account for acid-base behaviors in non-aqueous solvents or reactions lacking proton involvement. By framing acids as electron-pair acceptors within his valence theory, Lewis aimed to unify diverse chemical interactions under a single electronic paradigm, applicable beyond aqueous environments. Early illustrations in Lewis's work included the use of the term "" for the resulting complexes, with the ammonia-boron trifluoride (NH₃·BF₃)—where the nitrogen from coordinates with the electron-deficient boron atom in BF₃—serving as a classic later example of such coordination. This highlighted how Lewis viewed such as direct manifestations of , with acids serving as electron-pair acceptors to form stable shared-pair bonds.

Evolution and comparisons to other theories

Following Gilbert N. Lewis's initial formulation in 1923, subsequent developments expanded the electron-pair donor-acceptor framework, with Mikhail Usanovich proposing a further in 1939 that incorporated electron transfer processes encompassing oxidation-reduction reactions. In Usanovich's view, an acid is any species that accepts electrons or anions (or donates cations), while a base donates electrons or anions (or accepts cations), thereby broadening the scope to include redox phenomena alongside donation-acceptance. This approach, published in the Journal of of the USSR, built on Lewis's emphasis on electron involvement by integrating it with electrochemical contexts, influencing later unified theories. In 1938, Lewis himself reformulated and extended his specifically for applications in , emphasizing practical examples of electron-pair acceptance in reactions involving non-protic species like compounds. This iteration, detailed in his Journal of the Franklin Institute article, clarified the role of Lewis acids in catalyzing organic transformations by forming temporary adducts, such as BF₃ with ethers, thereby bridging theoretical concepts with synthetic utility. Compared to the contemporaneous Brønsted-Lowry (1923), which defines acids as proton donors and bases as proton acceptors, the Lewis framework is more encompassing, as Brønsted-Lowry reactions represent a subset where the proton (H⁺) acts as the Lewis acid and the base accepts the from its conjugate. For instance, BF₃ qualifies as a Lewis acid without involving protons, highlighting the limitations of the protic-focused Brønsted-Lowry model in non-aqueous or aprotic environments. The theory evolved further in the 1960s through Ralph G. Pearson's hard-soft acid-base (HSAB) principle, which built directly on Lewis's donor-acceptor paradigm by classifying acids and bases according to their and preferences. Pearson's seminal 1963 paper in the Journal of the posited that hard acids prefer hard bases (low , high ) and soft acids prefer soft bases (high , low ), providing a predictive tool for stability without altering the core electron-pair mechanism. In the , the acid-base concept has been unified with frontier (FMO) theory, originally advanced by Klopman in 1968, through computational and conceptual (DFT) frameworks that quantify interactions via highest occupied (HOMO) and lowest unoccupied (LUMO) overlaps. Recent reviews, such as the 2023 analysis completing the Klopman-Salem model, incorporate density polarization effects to refine predictions of Lewis adduct formation energies, demonstrating how FMO descriptors enhance understanding of reactivity in complex systems like organometallic . This integration resolves earlier ambiguities by treating acid-base reactions as charge-transfer processes modulated by orbital symmetries, with applications in modern quantum chemical simulations.

References

  1. [1]
    The Lewis Definitions of Acids and Bases
    In the Lewis theory of acid-base reactions, bases donate pairs of electrons and acids accept pairs of electrons.
  2. [2]
    16.2 Lewis Acids and Bases – Chemistry Fundamentals
    A Lewis acid is a species that can accept an electron pair, whereas a Lewis base has an electron pair available for donation to a Lewis acid. Complex ions are ...
  3. [3]
    Lewis Acids/Bases; HSAB Theory
    According to this theory, a base is defined as an electron pair donor, and an acid as an electron pair acceptor. Donation of an electron pair from base to acid ...
  4. [4]
    [PDF] INTRODUCTION TO LEWIS ACID-BASE CHEMISTRY
    Lewis acids and bases are defined in terms of electron pair transfers. A Lewis base is an electron pair donor, and a Lewis acid is an electron pair acceptor. ...
  5. [5]
    15.2 Lewis Acids and Bases - Chemistry 2e | OpenStax
    Feb 14, 2019 · In 1923, G. N. Lewis proposed a generalized definition of acid-base behavior in which acids and bases are identified by their ability to ...<|control11|><|separator|>
  6. [6]
    Lewis Acids and Bases | Chemistry: Atoms First - Lumen Learning
    A Lewis acid-base reaction occurs when a base donates a pair of electrons to an acid. A Lewis acid-base adduct, a compound that contains a coordinate covalent ...Missing: process | Show results with:process
  7. [7]
    Quantum Chemistry–Machine Learning Approach for Predicting ...
    May 19, 2023 · ... Lewis adduct clusters, whose structural motifs are absent in the training chemical space. ... for Me3N-AlCl3 (D0(298) = 49.3 kcal/mol). A strong ...
  8. [8]
    An easy-to-perform evaluation of steric properties of Lewis acids
    Mar 1, 2023 · Steric interactions influence the shape and the reactivity of molecules such as Lewis acids and Lewis bases. As steric effects have a decisive ...Missing: solvent | Show results with:solvent
  9. [9]
    Probing the Impact of Solvent on the Strength of Lewis Acids via ...
    Feb 9, 2023 · We investigate the impact of solvent effects on Lewis acids for the first time as measured by the fluorescent Lewis adduct (FLA) method.
  10. [10]
    Lewis Concept of Acids and Bases - Chemistry LibreTexts
    Jan 29, 2023 · The aluminum ion is the metal and is a cation with an unfilled valence shell, and it is a Lewis Acid. Water has lone-pair electrons and is an ...Lewis Acids · Lewis Bases · Complex Ion / Coordination... · Amphoterism
  11. [11]
    Lewis Acids And Bases - BYJU'S
    Jun 18, 2020 · Examples of Lewis Acids · H+ ions (or protons) can be considered as Lewis acids along with onium ions like H3O+. · The cations of d block elements ...Lewis Acid · Examples of Lewis Acids · Lewis Base · Examples of Lewis Bases
  12. [12]
    Electrostatics Explains the Reverse Lewis Acidity of BH3 and Boron ...
    Sep 22, 2021 · However, the acidity of boron trihalides unexpectedly follows the opposite trend: BF3 < BCl3 < BBr3. The most accepted explanation for this ...
  13. [13]
    Which is more acidic BF3 or AlCl3? - Chemistry Stack Exchange
    Jan 24, 2018 · Since Lewis acids are electron acceptors, BF3 is a stronger Lewis acid than AlCl3. Thus BF3 is better Lewis acid than AlCl3.
  14. [14]
    Review Applications of metal complexes in analytical chemistry
    Feb 15, 2024 · In this chemical interaction, the central metal element acts as the Lewis acid, accepting electron pairs from the ligands [1]. Ligands can ...
  15. [15]
    [PDF] Transition Metal Carbonyls
    CO is classed as a soft ligand because it is capable of accepting metal dπ electrons by back bonding, i.e. it is a σ-donor π-acceptor ligand. Metal Carbonyls: ...
  16. [16]
    Main Group Lewis Acid-Mediated Transformations of Transition ...
    May 10, 2016 · The bulk of this Review focuses on reactions where the transition metal and Lewis acid act in cooperation, and includes sections on carbonyl ...Missing: acidity | Show results with:acidity
  17. [17]
    Frustrated Lewis pairs reactivity across the periodic table: The 2024 ...
    This article presents an overview of some of our developments of the chemistry of “frustrated Lewis pairs” (FLPs). We begin with a brief outline of the ...
  18. [18]
  19. [19]
    Lewis Acids and Bases - Oregon State University
    A bare proton carries a positive charge and no electrons--the perfect Lewis acid. It will accept an electron pair from wherever it can find it to fill its 1s ...
  20. [20]
    [PDF] Chemical Bonding Michael Morse, University of Utah morse@chem ...
    Are there acids and bases that don't involve proton transfer? YES! And that's where things get interesting. Page 4. Lewis Acids and Bases: G. N. Lewis (1875- ...
  21. [21]
    Cationic Lewis Acids
    Cationic Lewis Acids. Metal cations have two characteristics that allow them to act as Lewis acids: 1. Their positive charge attracts electrons.
  22. [22]
    The Grignard Reagents | Organometallics - ACS Publications
    The Grignard reagents probably have been the most widely used organometallic reagents. Most of them are easily prepared in ethereal solution.
  23. [23]
    The Influence of Main Group Metallic Lewis Acids on the Formation ...
    Aug 6, 2025 · This mini-review provides a concise overview mainly on the behavior of main group metallic salts derived from lithium, magnesium and aluminum.Missing: Li+ Mg2+
  24. [24]
    The Hydration and Hydrolysis of Metal Cations
    The hydration energy of a cation depends on the charge and radius of the cation. Hydration energy also depends on the electronegativity of the element.
  25. [25]
    [PDF] Coordination Chemistry I: Structures and Isomers
    Nov 13, 2015 · Ambidentate ligands can bind to a metal using either of several atoms. isothiocyanate binds through N thiocyanate binds through S nitro NO.
  26. [26]
    Coordination Complexes and Ligands
    They are also known as complex ions or coordination complexes because they are Lewis acid-base complexes. ... Ethylenediamine (en) is a typical bidentate ligand.
  27. [27]
    CHM 501 Lecture 12 Introduction to Transition Metal Chemistry
    Ligands. Usually ligands are Lewis bases (nonmetals, either anions or neutrals with lone pairs) but this is not required. ambidentate : two available binding ...
  28. [28]
    Coordination Compounds
    Recall that the metal–ligand interaction is an example of a Lewis acid–base interaction. ... For example, the complex of Ni 2+ with three ethylenediamine ligands, ...
  29. [29]
    [PDF] Bridging cyanides from cyanoiron metalloligands to redox-active ...
    Cyanide, as an ambidentate ligand, plays a pivotal role in providing a simple diatomic building-block motif for controlled metal aggregation (M–CN–M').
  30. [30]
    Detection of Ethylene with Defined Metal Complexes - NIH
    The model for describing this dual mode of metal-ligand bonding is known as the Dewar-Chatt-Duncanson model. As a result, it is not surprising that transition ...
  31. [31]
    Interplay between Gold(I)-Ligand Bond Components and Hydrogen ...
    Jan 16, 2019 · This methodology provides the quantification of the Dewar–Chatt–Duncanson (DCD) components for the Au–C and Au–Se bonds in the absence and presence of the ...
  32. [32]
    [PDF] The Reactivity of Ni Complexes in Conjugated Polymer Synthesis by ...
    We found that π-binding energies followed the Dewar-Chatt-Duncanson model, indicating that Ni π-complexes are more strongly bound than Pd π-complexes, and π- ...
  33. [33]
    and Heteropolymetallic Complexes of the Hybrid, Ambidentate N ...
    Nov 15, 2018 · We have conducted a complete exploration of the coordination chemistry of the hybrid ambidentate IMes-acac ligand 1–, obtained by fusing a ...
  34. [34]
    Luminescent First-Row Transition Metal Complexes | JACS Au
    Sep 24, 2021 · Nature Chemistry (2010) ... These efforts involve new synthetic directions, including the choice of metal as well as new ambidentate ligands for ...Figure 4 · Figure 5 · Figure 6Missing: post- | Show results with:post-<|control11|><|separator|>
  35. [35]
    Future Energy The Power of Hydrides - ScienceDirect.com
    Apr 15, 2020 · As a result, H− is recognized as a soft but strong Lewis base as well as a potent electron donor, which differs distinctly from its neutral ...
  36. [36]
    Strong Bases Design: Predicted Limits of Basicity - ACS Publications
    In this article, the reasons for the instability of very basic compounds are investigated by means of high-level quantum-chemical calculations.
  37. [37]
    Lewis Adduct - an overview | ScienceDirect Topics
    A Lewis adduct is defined as a complex formed when a Lewis acid, which accepts electron pairs, interacts with a Lewis base, which donates electron pairs, ...
  38. [38]
    Quantitative evaluation of the HSAB [hard-soft acid-base] concept
    On the applicability of the HSAB principle through the use of improved computational schemes for chemical hardness evaluation. Journal of Computational ...Missing: validation | Show results with:validation
  39. [39]
    Chemical hardness-driven interpretable machine learning approach ...
    Dec 9, 2021 · The interpretable ML regression, classification, and data analysis lead to a new hypothesis that the highly stable 2DO materials follow the HSAB ...
  40. [40]
    Quantum Chemical Cluster Model and HSAB Approaches - MDPI
    Jan 16, 2023 · In the present study, the interaction between organophosphorus pesticides and cholinesterase enzymes was investigated by quantum chemical cluster model and ...
  41. [41]
    Quantification of Lewis Acidity and Lewis Basicity: A Density‐Based ...
    Aug 11, 2025 · This work suggests that density-based frameworks could provide a powerful and novel foundation for understanding the hard and soft acids and ...
  42. [42]
    [PDF] Determination of Lewis Acidity using 31P NMR - Carbon
    31P NMR uses triethylphosphine oxide to measure Lewis acidity by correlating the chemical shift with the 'Acceptor Number' (Gutmann-Beckett method).
  43. [43]
    Gutmann acceptor and donor numbers - Stenutz
    The Gutmann Acceptor- (AN) and Donor Number (DN) are measures of the strength of solvents as Lewis acids or bases.
  44. [44]
    An easy-to-perform evaluation of steric properties of Lewis acids
    Feb 6, 2023 · The most wide-spread measure to evaluate the strength of a Lewis acid nowadays is probably fluoride ion affinity (FIA), which is defined as the ...
  45. [45]
    Analyzing Fluoride Binding by Group 15 Lewis Acids: Pnictogen ...
    Aug 8, 2023 · The computed fluoride ion affinity (FIA) is a valuable descriptor to assess the Lewis acidity of a compd. Despite its widespread use, the ...Introduction · Computational Methods · Results and Discussion · Conclusions
  46. [46]
    The global electrophilicity index as a metric for Lewis acidity
    The global electrophilicity index, GEI, is used as a general, quantitative and base-independent metric of Lewis acidity.Missing: ω DFT Parr
  47. [47]
    Scales of Lewis Basicities toward C-Centered Lewis Acids ...
    Lewis Acidity Scale of Diaryliodonium Ions toward Oxygen, Nitrogen, and Halogen Lewis Bases. Journal of the American Chemical Society 2020, 142 (11) , 5221 ...
  48. [48]
    [PDF] Evaluated Gas Phase Basicities and Proton Affinities of Molecules
    Oct 15, 2009 · Tables giving the molecules ordered (1) according to proton affinity and (2) according to empirical formula, sorted alphabetically are provided.
  49. [49]
    Lewis Acids: From Conventional Homogeneous to Green ...
    Journal of the American Chemical Society 2025, 147 (4) , 3315-3332. https://doi.org/10.1021/jacs.4c13481. Lydia A. Dunaway, Audrey G. Davis, Victoria J ...
  50. [50]
    Bifunctional Asymmetric Catalysis: Cooperative Lewis Acid ... - NIH
    Bifunctional asymmetric catalysis uses cooperative Lewis acid/base systems to activate both electrophilic and nucleophilic reaction partners, often in a ...
  51. [51]
    Lewis Acid-Catalyzed Enantioselective Friedel–Crafts Alkylation of ...
    May 11, 2020 · The Friedel–Crafts alkylation products can be obtained in high yields and excellent enantioselectivities.
  52. [52]
    Zinc(II) Carboxylate Coordination Polymers with Versatile Applications
    The ability of Zn(II) to act as a good Lewis acid allowed the involvement of Zn-CBCPs either in dye elimination from wastewater through photocatalysis or in ...
  53. [53]
    Zeolites in Adsorption Processes: State of the Art and Future Prospects
    Oct 19, 2022 · Zeolites are crystalline microporous aluminosilicates widely used as catalysts, adsorbents, and ion exchangers. They belong to the tectosilicate ...
  54. [54]
    Dynamical Equilibrium Brønsted & Lewis Sites in Zeolites
    Jun 7, 2023 · The structure and properties of Brønsted acid sites in zeolite are well understood, while those of Lewis acid sites in zeolites are not fully ...
  55. [55]
    Lewis base anchored DhaTph covalent-organic framework for ...
    Jan 1, 2024 · DhaTph can serve as a potential adsorbent for selective capture of Th(IV) ion, and provides a guideline for designing effective COFs-based adsorbents.
  56. [56]
    Covalent Organic Frameworks: From Structures to Applications - MDPI
    The pyridyl groups are the Lewis base for catalyzing the reaction between malanonitrile and an aromatic aldehyde with a yield of 72%. Polymers 15 01279 sch004.
  57. [57]
    Interconvertible and rejuvenated Lewis acidic electrolyte additive for ...
    Jul 24, 2025 · In this work, we report a handy yet effective electrolyte additive of Lewis acidic multivalent metal cations, which enables low E/S ratio ...
  58. [58]
    Lewis Acid-Triggered In-Situ Gelation of Ether Electrolytes for ...
    Oct 20, 2025 · Highlights. The Lewis acid facilitated the in-situ ring-opening polymerization of the electrolyte. The Sb3+ achieved in-situ alloying of the ...
  59. [59]
    Valence and The Structure of Atoms and Molecules - Internet Archive
    Jun 23, 2024 · Valence and The Structure of Atoms and Molecules. by: Gilbert Newton Lewis ... PDF download · download 1 file · PNG download · download 1 file.
  60. [60]
    The Usanovich definition of acids and bases – A forgotten pioneer of ...
    Oct 8, 2020 · This article portrays not only the Usanovich definition itself, but also demonstrates its value as a historical context for teaching nature ...
  61. [61]
    (PDF) Usanovich and Nernst colliding: inconsistencies in the all-in ...
    Oct 4, 2023 · Among the many acid-base concepts, the theory of Usanovich is one of the least known despite the most general scope including almost all ...
  62. [62]
    Acids and bases - ScienceDirect.com
    Journal of the Franklin Institute · Volume 226, Issue 3, September 1938, Pages 293-313. Journal of the Franklin Institute. Acids and bases☆. Author links open ...