Fact-checked by Grok 2 weeks ago

Electron-withdrawing group

An electron-withdrawing group (EWG) is a attached to an molecule that attracts toward itself from adjacent atoms or the molecular framework, primarily through inductive or mechanisms. Inductive withdrawal occurs via bonds due to the high of the group, polarizing bonds and depleting , while withdrawal involves delocalization through pi systems, often stabilizing nearby positive charges or destabilizing negative ones. These effects make EWGs crucial in modulating molecular properties and reactivity in . Common examples of EWGs include the nitro group (-NO₂), cyano group (-CN), carbonyl-containing groups such as acyl (-COR) and (-COOR), trifluoromethyl (-CF₃), and like (-F) or (-Cl). exhibit both inductive withdrawal (strong -I effect) and donation (+M), resulting in net deactivation but ortho/para direction in some contexts, whereas groups like -NO₂ are strong deactivators via both effects. The strength of withdrawal varies; for instance, -NO₂ is a potent -R group, while -CF₃ primarily acts inductively. In electrophilic aromatic substitution (EAS), EWGs deactivate the benzene ring by reducing electron density, slowing reaction rates—sometimes by factors exceeding 10 million for nitrobenzene compared to benzene—and direct incoming electrophiles to the meta position to avoid destabilizing the positively charged intermediate. They also enhance acidity of proximal functional groups, such as lowering the pKa of carboxylic acids through stabilization of the conjugate base; for example, fluoroacetic acid has a pKa of 2.7 versus 4.8 for acetic acid. Additionally, EWGs destabilize carbocations in solvolysis reactions, reducing Sₙ1 rates, and influence nucleophilic substitutions by activating sites for attack.

Fundamentals

Definition

An electron-withdrawing group (EWG) is a molecular entity that withdraws substantial amounts of from a neighboring molecular entity through inductive and/or effects. These groups typically involve atoms or functional units with high or multiple bonds that facilitate electron displacement away from a , thereby altering the electronic environment of the attached organic framework. The concept of electron-withdrawing groups emerged in the early amid studies on effects in derivatives, with Louis P. Hammett formalizing quantitative relationships in through an that quantified how such groups influence rates and equilibria. Hammett's work built on prior qualitative observations of how substituents modify the reactivity of aromatic systems by shifting , laying the groundwork for understanding EWGs as modulators of molecular electronics. In molecules, EWGs play a pivotal role by polarizing bonds and redistributing charge, which affects the overall in sigma frameworks or pi systems. This withdrawal establishes a foundational electronic gradient that can stabilize adjacent negative charges or destabilize positive ones, influencing the intrinsic properties of the without altering its core .

Mechanisms of Electron Withdrawal

-withdrawing groups (EWGs) primarily exert their influence through two distinct mechanisms: the and the resonance . The involves the permanent polarization of sigma bonds due to differences in between atoms, leading to withdrawal across single bonds. This results in a decrease in at the reaction center, denoted as the -I . In the inductive mechanism, an atom or group with higher attracts electrons from adjacent bonds, creating a partial positive charge that propagates through the framework. This stabilizes nearby electron-deficient sites, such as carbocations, or destabilizes electron-rich ones. The effect is transmitted via the overlap of orbitals along the molecular chain, without requiring conjugation. The effect, in contrast, operates through delocalization of pi electrons in conjugated systems, allowing EWGs to withdraw via pi-orbital overlap. This mechanism, often denoted as the -M or -R effect, stabilizes adjacent carbanions by dispersing negative charge or enhances cations by accepting electron density into the group's . Resonance withdrawal is particularly pronounced in meta-directing groups, where the substituent's empty orbitals or electronegative atoms interact with the pi cloud. Many EWGs exhibit hybrid behavior, combining both inductive and resonance effects, with the dominant mode depending on the group's structure and position. For instance, the inductive component relies on orbital overlap for through-bond transmission, while involves lateral p-orbital interactions in pi-conjugated frameworks, as illustrated by the differing maps in representations. This interplay can lead to net electron withdrawal even when individual effects oppose each other. The strength of electron withdrawal is influenced by several factors, including the distance from the reaction center, where the inductive effect diminishes rapidly—typically becoming negligible beyond three carbon atoms—due to the exponential decay of polarization through sigma bonds. Electronegativity differences, quantified on the Pauling scale (ranging from about 0.7 for alkali metals to 4.0 for ), determine the initial polarization magnitude, with more electronegative atoms enhancing withdrawal. Solvent effects modulate this polarization; in polar media, the dielectric constant can either amplify or attenuate the inductive field by solvating charges, thereby altering the effective transmission of ./01%3A_Map-Inorganic_Chemistry-I(LibreTexts)/03%3A_Simple_Bonding_Theory/3.02%3A_Valence_Shell_Electron-Pair_Repulsion/3.2.03%3A_Electronegativity_and_Atomic_Size_Effects)

Examples

Inductive Electron-Withdrawing Groups

Halogens such as , , , and iodine serve as classic examples of groups that withdraw electrons primarily through the . Their high polarize the carbon-halogen bond, placing a partial positive charge (δ⁺) on the carbon atom adjacent to the halogen and a partial negative charge (δ⁻) on the halogen itself. This polarization arises from the unequal sharing of electrons in the σ-bond, with the effect strongest for due to its highest electronegativity and decreasing down the group to iodine as atomic size increases and electronegativity diminishes. The trifluoromethyl group (-CF₃) exemplifies a highly effective inductive electron-withdrawing , where the carbon is bonded to three atoms. The cumulative of these fluorines intensifies the withdrawal of through the σ-bonds, creating a significant partial positive charge on the attached carbon and enhancing electrophilicity at nearby sites. This makes -CF₃ particularly potent in modulating molecular reactivity via inductive means. Alkyl ammonium groups, represented as -NR₃⁺ (where R denotes alkyl substituents), demonstrate strong inductive electron withdrawal attributed to the positive charge on the , which electrostatically attracts across intervening σ-bonds. This cationic character renders the group one of the most powerful inductive withdrawers, far exceeding neutral substituents. Such groups find practical application in , where the enables the solubilization and transport of anionic species across phase boundaries in biphasic reaction systems. A key characteristic of inductive electron withdrawal is its rapid attenuation with increasing distance from the substituent, which confines the influence to short-range interactions within a few bonds. This distance-dependent decay underscores the σ-bond-mediated nature of the effect, distinguishing it from longer-range delocalized mechanisms.

Resonance Electron-Withdrawing Groups

Resonance electron-withdrawing groups (EWGs) primarily deplete electron density from an adjacent through conjugation, involving the overlap of p-orbitals that allows delocalization of electrons away from the core structure. These groups are characterized by their ability to stabilize negative charges or electron-rich intermediates via donation from the pi system to the substituent, contrasting with purely sigma-based inductive effects. The nitro group (-NO₂) exemplifies a strong resonance EWG, where the nitrogen atom connects to the pi system, enabling electron delocalization toward the electronegative oxygen atoms. Resonance structures depict the aromatic ring's pi electrons contributing to a quinoid form, with the negative charge residing on the oxygens, thereby withdrawing density from the ring. This delocalization makes the nitro group particularly effective in conjugated systems, such as , where it influences reactivity across multiple bonds. Carbonyl-containing groups, including aldehydes (-CHO), ketones (-COR), carboxylic acids (-COOH), and esters (-COOR), function as resonance EWGs through their conjugated pi bonds. The carbonyl's C=O accepts electrons from an attached , forming hybrids where the carbon bears a partial positive charge and the oxygen a negative one, pulling density away from the . For instance, in , the aldehydic carbonyl conjugates with the aromatic ring, delocalizing pi electrons into the C=O antibonding orbital. Similar occurs in carboxylic acids and esters, where the oxygen atoms enhance the withdrawal, though the effect is modulated by the specific substituents. The cyano group (-CN) operates via resonance withdrawal through its linear pi bonding, where the triple bond between carbon and nitrogen allows electrons from an adjacent pi system to delocalize toward the electronegative nitrogen. Resonance forms show the substrate's pi density contributing to a structure with positive charge on the attachment carbon and negative on nitrogen, effectively reducing electron availability in the core. This makes the cyano group a potent deactivator in aromatic systems, as seen in benzonitrile. Unlike inductive effects, which diminish rapidly with distance due to sigma bond transmission, resonance effects in these EWGs extend over longer ranges through conjugated pi networks, often resulting in meta-directing behavior in by destabilizing and positions via .

Effects on Acidity

Brønsted-Lowry Acidity

Electron-withdrawing groups (EWGs) enhance Brønsted-Lowry acidity by stabilizing the conjugate base through inductive and resonance effects, which disperse the negative charge developed upon proton loss and thereby lower the pKa value. In the inductive effect, EWGs withdraw electron density via sigma bonds, while in the resonance effect, they delocalize the charge through pi systems, particularly when the EWG is conjugated with the site of deprotonation. This stabilization makes proton donation more favorable, as the energy difference between the acid and its conjugate base decreases. For example, nitroacetic acid, with its strongly electron-withdrawing nitro group at the alpha position, has a pKa of approximately 1.7, compared to 4.76 for unsubstituted acetic acid, illustrating the profound impact of such groups on acidity. The provides a quantitative framework for assessing EWG effects on acidity in aromatic systems, expressed as: \log\left(\frac{K}{K_0}\right) = \rho \sigma where K and K_0 are the acid dissociation constants of the substituted and parent s, respectively, \sigma is the constant reflecting the electron-withdrawing ability of the group, and \rho is the reaction constant (set to 1 for ionization at 25°C). Positive \sigma values for EWGs predict increased acidity. In the series, the para-nitro has \sigma = 0.78, leading to a of 3.44 for p-nitrobenzoic acid versus 4.20 for ; this \DeltapKa of 0.76 units aligns closely with the \sigma value, confirming the nitro group's strong and inductive withdrawal of electrons to stabilize the anion. A clear demonstration of cumulative inductive effects from multiple EWGs is seen in halogenated carboxylic acids, where each additional electronegative atom further stabilizes the conjugate base. , bearing three atoms at the alpha carbon, exhibits a of 0.66—substantially lower than the 4.76 of acetic acid—due to the combined electron-withdrawing influence of the chlorines, which pull away from the group through sigma bonds. Alpha-substituents in carboxylic acids primarily exert their influence via inductive stabilization of the conjugate base, as the proximity allows direct sigma-bond transmission of electron withdrawal without significant involvement. For instance, introducing a single at the alpha position in reduces the to 2.86 from 4.76 for acetic acid, a shift of approximately 1.9 units attributable to the of dispersing the negative charge on the . This effect intensifies with multiple alpha-halogens, underscoring the role of inductive forces in modulating acidity for aliphatic systems.

Lewis Acidity

Electron-withdrawing groups (EWGs) enhance the Lewis acidity of a by reducing the on the central atom, thereby increasing its ability to accept an from a Lewis base. This effect is primarily inductive for groups like , where the high pulls electrons through sigma bonds, rendering the central atom more electrophilic. For instance, in (BF₃), the three atoms withdraw from the center via inductive effects, making BF₃ a potent Lewis capable of coordinating with bases such as ethers or amines. In contrast, trimethylborane (BMe₃) exhibits much weaker Lewis acidity because the methyl groups act as electron donors, increasing the electron density on and reducing its electrophilicity. The influence of EWGs on Lewis acidity can be quantified using the Tolman electronic parameter (TEP), which measures the electron-withdrawing ability of ligands through the A₁ symmetric CO stretching frequency (ν(CO)) in Ni(CO)₃L complexes. A higher ν(CO) indicates greater electron withdrawal by the ligand L, as it reduces back-donation from the metal to , strengthening the bond. For example, (PF₃), with its electronegative substituents, shows a ν(CO) of 2110.8 cm⁻¹, reflecting strong electron withdrawal and enhanced Lewis acidity at the phosphorus center compared to trimethylphosphine (PMe₃), which has a ν(CO) of 2056.1 cm⁻¹ due to the donating methyl groups. This parameter highlights how inductive EWGs like amplify the acceptor properties of ligands in coordination chemistry. In coordination chemistry and , molecules bearing EWGs often serve as effective acids by providing empty orbitals for acceptance. Acyl chlorides (RCOCl) exemplify this, where the atom inductively withdraws electrons from the carbonyl carbon, increasing its electrophilicity and enabling coordination with bases in reactions such as nucleophilic acyl substitutions. Similarly, tris(trifluoromethyl)borane () demonstrates extreme acidity due to the strongly withdrawing -CF₃ groups, which render the boron center highly receptive to nucleophiles; this compound is used in for activating inert bonds. A notable consequence is the enhanced affinity for ions, as seen in B(CF₃)₃, where computational affinities reveal a stronger F⁻ interaction (approximately 555 kJ/mol) compared to BF₃ (350 kJ/mol), underscoring the amplifying role of perfluoroalkyl EWGs in stabilizing adducts.

Effects on Reactivity

Electrophilic Aromatic Substitution

Electron-withdrawing groups (EWGs) deactivate aromatic rings toward (EAS) by reducing the in the π-system through inductive and effects, thereby increasing the energy barrier for the rate-determining addition of the to form the Wheland . This deactivation is particularly pronounced for strong EWGs like the nitro group; for instance, the of proceeds approximately 106 times slower than that of due to the diminished nucleophilicity of the ring./16%3A_Chemistry_of_Benzene_-_Electrophilic_Aromatic_Substitution/16.04%3A_Substituent_Effects_in_Electrophilic_Substitutions) In contrast, EWGs such as -NO2 exert their influence primarily through delocalization of electrons away from the ring. The meta-directing effect of EWGs arises because attack at the or positions leads to Wheland intermediates where a key structure places positive charge directly adjacent to the electron-withdrawing , causing significant destabilization; in the Wheland intermediate, however, all forms distribute the positive charge without such adjacency, making meta attack relatively more favorable. For , yields 93% meta product (with only 6% and 1% ), illustrating this selectivity as the meta position experiences the least electron withdrawal in the ./16%3A_Chemistry_of_Benzene_-_Electrophilic_Aromatic_Substitution/16.04%3A_Substituent_Effects_in_Electrophilic_Substitutions) In compounds bearing multiple substituents, the directing effects can interact; for example, chlorobenzene, where the halogen is a weakly deactivating ortho/para director, undergoes nitration to give 35% ortho and 64% para isomers (with negligible meta product), but slower than benzene by a factor of about 30 due to partial electron withdrawal. However, in nitrochlorobenzene (e.g., 1-chloro-4-nitrobenzene), the strongly meta-directing nitro group dominates, directing further electrophilic substitution predominantly to the meta position relative to itself, overriding the halogen's influence and resulting in over 90% meta selectivity in reactions like halogenation./16%3A_Electrophilic_Aromatic_Substitution/16.13%3A_Electrophilic_Aromatic_Substitution_of_Substituted_Benzenes) When multiple EWGs are present, they can dramatically alter reactivity beyond EAS, enabling (SNAr); for instance, in 1-chloro-2,4-dinitrobenzene, the two groups cumulatively withdraw electrons to stabilize the anionic formed upon , facilitating displacement of chloride by nucleophiles like amines or alkoxides under mild conditions. This contrasts with unactivated aryl halides, which resist SNAr without such stabilization.

Redox Potentials

Electron-withdrawing groups (EWGs) affect potentials by delocalizing from the redox-active center, thereby stabilizing the oxidized form relative to the reduced form. This stabilization occurs primarily through inductive and effects, which lower the energy of the lowest unoccupied (LUMO) in reductions or raise the highest occupied (HOMO) energy barrier in oxidations. Consequently, the standard (E°) shifts positively, making reduction of the oxidized easier and oxidation of the reduced more difficult. For instance, in one-electron processes, EWGs enhance the oxidizing strength of the couple by approximately 200–600 mV depending on the system and . A prominent illustration is the ferrocene/ferrocenium (Fc/Fc⁺) redox couple, where the parent exhibits E° ≈ +0.40 V vs. in . Substitution with an (-COCH₃), a resonance EWG, shifts this potential anodically to ≈ +0.67 V vs. , representing a 270 mV increase due to the carbonyl's ability to conjugate with and delocalize positive charge on the ferrocenium ion. Similarly, in quinone chemistry, p-benzoquinone displays a first one-electron reduction potential (Q/Q⁻) of -401 mV vs. in aprotic media, while nitro-substituted analogs like 2,5-dinitro-1,4-benzoquinone show a markedly more positive value near +199 mV vs. for related dinitro derivatives, amplifying their role as potent oxidants in electron-transfer reactions. Cyclic voltammetry provides direct evidence of these effects through observed shifts in peak potentials. EWGs cause anodic shifts in oxidation peak potentials (E_{pa}) by stabilizing intermediate cations, often converting irreversible processes into quasi-reversible ones with smaller peak separations (ΔE_p < 100 mV). This stabilization mitigates rapid follow-up reactions, such as dimerization, allowing cleaner electrochemical characterization and control in synthetic electrochemistry. In organometallic complexes, EWGs like ester groups (-COOR) on cyclopentadienyl ligands elevate metal-centered redox potentials, facilitating selective oxidations. For example, -COOR-substituted ferrocenes exhibit potentials raised by 200–400 mV relative to unsubstituted analogs, enabling precise matching to mild oxidants like ceric ammonium nitrate for targeted functionalization in multistep syntheses without over-oxidation of sensitive moieties.

Comparisons and Quantification

Comparison with Electron-Donating Groups

Electron-withdrawing groups (EWGs) exert an inductive (-I) and/or resonance (-M) effect that pulls electron density away from the attached atom or system, whereas (EDGs) push electron density toward it via inductive (+I) and/or resonance (+M or +R) effects. This opposition results in contrasting charge stabilization: EWGs stabilize adjacent anions by dispersing negative charge, enhancing acidity, while EDGs stabilize cations by increasing electron density at the site. In terms of reactivity, particularly electrophilic aromatic substitution (EAS), EWGs deactivate the aromatic ring and direct incoming electrophiles to the meta position by withdrawing electron density, making ortho and para positions less favorable for the positively charged intermediate. In contrast, EDGs activate the ring and direct ortho/para by donating electron density, stabilizing the intermediate at those positions. The following table illustrates this with representative examples:
Group TypeExample GroupExample CompoundRing EffectDirecting Effect
EWG-NO₂NitrobenzeneDeactivatingMeta
EDG-NH₂AnilineActivatingOrtho/Para
For acidity, the nitro group (-NO₂) as an EWG increases the acidity of nearby protons, such as in substituted benzoic acids, by stabilizing the conjugate base through electron withdrawal. Conversely, the hydroxy group (-OH) as an EDG decreases acidity in similar contexts by donating electron density, destabilizing the conjugate base. Halogens represent ambiguous substituents, acting as inductive EWGs (-I) due to their electronegativity but as resonance EDGs (+R) through lone pair donation into the π system. The net effect is deactivation of the ring overall, yet ortho/para direction in EAS because the resonance donation outweighs inductive withdrawal in stabilizing the key intermediate.

Hammett Sigma Constants

The Hammett equation, formulated by Louis P. Hammett in 1937, quantifies the influence of substituents on the rates and equilibria of reactions involving benzene derivatives through a linear free-energy relationship. It is given by \log \left( \frac{K}{K_0} \right) = \rho \sigma where K and K_0 are the equilibrium or rate constants for the substituted and unsubstituted (parent) compounds, respectively; \sigma is the substituent constant measuring the electronic effect of the group; and \rho is the reaction constant reflecting the sensitivity of the process to these effects. For electron-withdrawing groups (EWGs), \sigma > 0, as they destabilize at the reaction center, accelerating reactions that build negative charge or decelerating those that build positive charge. Substituent constants \sigma are position-dependent, with distinct values for meta (\sigma_m) and para (\sigma_p) placements on the ring; meta substitution primarily captures inductive () effects, while para includes both inductive and contributions, allowing differentiation between these mechanisms. This distinction evolved from Hammett's initial work, which emphasized empirical correlations for aromatic systems and later refinements to account for dominance in para positions. Representative \sigma values for common EWGs, derived from the ionization of substituted benzoic acids in at 25°C, are listed below; higher positive values denote stronger withdrawal.
Substituent\sigma_m\sigma_p
-NO₂0.710.78
-CN0.560.66
-CF₃0.430.54
-CHO0.350.42
-CO₂H0.370.45
-F0.340.06
These values, compiled from extensive tabulations, illustrate the group's exceptional withdrawing power via in the para position, contrasted with ' weaker para effects due to partial donation. The equation's \rho parameter is normalized to 1.00 for the ionization of benzoic acids, serving as the reference for defining \sigma, but varies with reaction type—positive \rho > 1 for processes sensitive to withdrawal, such as base-catalyzed reactions. While highly effective for meta- and para-substituted benzenes, the Hammett approach has limitations in ortho-substituted or non-aromatic systems, where steric effects or direct conjugation interfere; extensions like Taft's polar substituent constants (\sigma^*) address aliphatic and ortho-benzoate systems by separating polar and steric influences.

References

  1. [1]
    None
    ### Summary of Electron-Withdrawing Groups (EWG) from Chapter 14
  2. [2]
  3. [3]
  4. [4]
    The Effect of Structure upon the Reactions of Organic Compounds ...
    Research Article January 1, 1937. The ... Quantum Alchemy Based Bonding Trends and Their Link to Hammett's Equation and Pauling's Electronegativity Model.Missing: original | Show results with:original<|control11|><|separator|>
  5. [5]
    [PDF] Hammett Deserved a Nobel Prize* - eScholarship
    Currently the major utility of the Hammett equation is for drawing mechanistic inferences. If is > 0, the reaction is accelerated by electron-withdrawing ...
  6. [6]
    Inductive and Resonance (Mesomeric) Effects - Chemistry Steps
    For electron-donating groups, a positive sign is used (+I, +M), and for electron-withdrawing groups, a negative sign is used (–I, –M). The same group may have ...
  7. [7]
  8. [8]
    Substituent Effects from the Point of View of Energetics and ...
    Jun 2, 2023 · However, some relations including them are hard to interpret due to the changes in the inductive/resonance character of the substituents.
  9. [9]
    Ch12: Substituent Effects - University of Calgary
    The resonance decreases the electron density at the ortho- and para- positions. Hence these sites are less nucleophilic, and so the system tends to react with ...
  10. [10]
    Inductive Effect | CK-12 Foundation
    Sep 30, 2025 · Electron-withdrawing groups (–I) increase acidity by stabilising the conjugate base, while electron-donating groups (+I) increase basicity by ...Missing: mechanism | Show results with:mechanism
  11. [11]
    Electron-Withdrawing and Donating Effects in Organic Chemistry
    Electron-withdrawing groups (EWGs) pull electrons away from a molecule or reaction center, decreasing electron density. Examples include -NO2, -CN, and -COOH.
  12. [12]
  13. [13]
    Superelectrophiles and the Effects of Trifluoromethyl Substituents
    ### Summary: CF3 as Inductive Electron-Withdrawing Group
  14. [14]
    Inductive electron-withdrawal from ammonium ion headgroups of ...
    The results indicate that the inductive effects of electron-withdrawing functionality diminish transfection activity in modest (2-4-fold) increments. The ...
  15. [15]
    [PDF] Phase-Transfer Catalysis (PTC) - Macmillan Group
    Apr 10, 2008 · As alkyl groups becomes larger and larger the rate of transfer ... □ Several reactions have been done using chiral quaternary ammonium salts as ...Missing: "-NR3+" electron withdrawing
  16. [16]
    Three-Step Method to Teach Inductive Effect - ACS Publications
    Jun 13, 2023 · This work developed a new method to teach inductive effect, named the “Three-step Method” (TsM): (1) establishing the concept of inductive effect.
  17. [17]
    [PDF] Aromatic Substitution 2 - UMD
    Feb 16, 2024 · The inductive effect is negative (-I) when the substituent is an electron withdrawing group ( halogen or adjacent positive charge). The ...
  18. [18]
    [PDF] resonance and induction tutorial
    For example, an oxygen atom in a hydroxy group (OH) is electron withdrawing by induction, but electron donating by resonance when placed in a position on the ...
  19. [19]
    Electrophilic Aromatic Substitution Reactions II - LON-CAPA OCHem
    Unlike inductive effects which are transmitted through s bonds, resonance effects involve non-bonding electrons and pi electrons.
  20. [20]
    [PDF] Acids and Bases - SDSU Chemistry
    Feb 3, 2019 · withdrawing. ❖ Inductive effects weaken as the distance from the group increases. • Inductive polarization of electron density transmitted ...Missing: longer | Show results with:longer
  21. [21]
    Acids and Bases – Organic Chemistry - Maricopa Open Digital Press
    The chlorine substituent can be referred to as an electron-withdrawing group because of the inductive effect (Figure ). Greater the electronegativity of the ...
  22. [22]
    [PDF] Sametz: CHEM 322 2010
    ... acid and of 2-nitroacetic acid. One has a. pKa of 1.7, and the other has a pKa of 4.8. Assign each structure with the correct pKa, and explain how you ...
  23. [23]
    [PDF] pka-compilation-williams.pdf - Organic Chemistry Data
    Apr 7, 2022 · pKa Data Compiled by R. Williams. pKa Values. INDEX. Inorganic. 2 ... Acetic acids, substituted. Phosphonates (Ref. 2). H-. 4.76*. 20. X. -H. -H.
  24. [24]
    Unit 4: Free Energy Relationships
    Such a substituent is considered to be an electron-withdrawing group (EWG), because electron density is increased at the reaction site in the product ...
  25. [25]
    [PDF] Hansch-et-al.-Table-of-Hammett-parameters.pdf - Wang Lab
    *Values in parentheses are suspected of being inaccurate. Calculated from eq 8. Calculated from eq 2. Substituents are arranged by molecular formula, C,H,; ...
  26. [26]
    4-Nitrobenzoic Acid | C7H5NO4 | CID 6108 - PubChem
    The pKa of 4-nitrobenzoic acid is 3.44(4), indicating that this compound will exist almost entirely in the anion form in the environment and anions generally do ...
  27. [27]
    Ionization Constants of Organic Acids - MSU chemistry
    trichloroacetic acid. CCl3CO2H · 0.23. 0.77 ; oxalic acid. (CO2H) · K1 = 6.5 * 10 · K2 = 6.1 * 10 · 1.2 4.2.
  28. [28]
    Appendix C: Dissociation Constants and pKa Values for Acids at 25°C
    Chapter 27 Appendix C: Dissociation Constants and pK a Values for Acids at 25°C ; Acetic acid, CH 3CO 2H · 1.75 × 10 ; Arsenic acid, H 3AsO · 5.5 × 10 ; Benzoic acid ...
  29. [29]
    [PDF] Chapter 19: Carboxylic Acids
    19.6: Substituents and Acid Strength. Substituents on the α-carbon influence the pKa of carboxylic acids largely through inductive effects.
  30. [30]
    Order of lewis acidity for BBr3, BCl3, BMe3, BH3, BPh3?
    Aug 1, 2018 · BBr3 is more acidic than BCl3 due to weaker pi back bonding in BBr3 as compared to BCl3.BCl3 is stronger acid than BH3 due to electron withdrawing effect of Cl.
  31. [31]
    Lewis Acids Catalyzed Annulations of Ynamides with Acyl Chlorides ...
    Jul 6, 2018 · A one-pot synthesis of 4-amino-2-naphthol derivatives is accomplished via a ZnI 2 -catalyzed tandem Friedel–Crafts reaction sequence.
  32. [32]
    Review Lewis acidity of boron compounds - ScienceDirect.com
    An analysis of the roles of different electronic and structural factors on Lewis acidity of boron compounds is presented in this review.
  33. [33]
    Developing simple boranes as phase transfer catalysts ... - ChemRxiv
    Oct 26, 2021 · Screening different boranes revealed a correlation between calculated fluoride affinity of the borane and nucleophilic fluorination reactivity, ...
  34. [34]
    Electronic effect of substituent groups on the oxidation and reduction ...
    Oct 15, 2023 · By introducing electron-donating or electron-withdrawing groups onto the tpy ligand, it is possible to adjust the redox characteristics of ...
  35. [35]
    Effect of Cycling Ion and Solvent on the Redox Chemistry of ...
    Jun 5, 2020 · This report summarizes our systematic investigation of the redox chemistry of a series of quinones with electron-withdrawing and electron-donating substituents.Introduction · Results and Discussion · Conclusion · Supporting Information
  36. [36]
    A New Approach to Infrared Spectroelectrochemistry Using a Fiber ...
    The oxidation of 1.0 mM acetylferrocene (2, E°' = 0.27 V vs ferrocene) was studied in CH2Cl2/0.1 M [NBu4][PF6] at 273 K by monitoring the position and ...
  37. [37]
  38. [38]
    Electrochemical Characterization of Aromatic Molecules with 1,4 ...
    Apr 12, 2021 · Electron-withdrawing groups such as in molecules 2 and 3 tended to increase stability and shift the redox potential to more positive values, ...
  39. [39]
    Chemical Redox Agents for Organometallic Chemistry
    The aim of this review is to show how one-electron oxidants and reductants have been used in preparative chemistry (incorporating both synthetic applications)Missing: EWGs | Show results with:EWGs
  40. [40]
    Carboxylic Acid Reactivity - MSU chemistry
    Electronegative substituents increase acidity by inductive electron withdrawal. As expected, the higher the electronegativity of the substituent the greater the ...
  41. [41]
    Typical Electron-Withdrawing Groups Are ortho, meta-Directors ...
    Apr 22, 2025 · Electron-withdrawing groups are traditionally considered meta-directing in aromatic substitution reactions.
  42. [42]
    Chapter 21:Reactions of Aromatics
    While the great majority of all substituents conduct themselves according to the rules, i.e., EWG's are meta-directing and rate-retarding, while EDG's are o,p- ...
  43. [43]
    Aromatic Reactivity - MSU chemistry
    Electrophilic nitration and Friedel-Crafts acylation reactions introduce deactivating, meta-directing substituents on an aromatic ring.Missing: EWG EDG<|control11|><|separator|>
  44. [44]
    [PDF] Ch17 Reactions of Aromatic Compounds
    1) Toluene reacts about 25 times faster than benzene under identical conditions. (We say toluene is activated toward electrophilic aromatic substitution, and ...
  45. [45]
    Electrophilic Aromatic Substitution AR5. Directing Effects - csbsju
    Substituents on the benzene ring also influence the regiochemistry of the reaction. That is, they control where the new substituent appears in the product.Missing: EWG EDG
  46. [46]
    [PDF] Carboxylic Acids and the Acidity of the O±H Bond 19±1
    A carboxylic acid is more acidic than an alcohol or phenol because its conjugate base is more effectively stabilized by resonance (19.9). OH. pKa 5. pKa = 16–18.
  47. [47]
    Substitution Reactions of Benzene and Other Aromatic Compounds
    This activation or deactivation of the benzene ring toward electrophilic substitution may be correlated with the electron donating or electron withdrawing ...
  48. [48]
    Hammett Sigma Constants* - Wired Chemist
    Hammett Sigma Constants* ; NO2**, 0.71, 0.78 ; NO, 0.62, 0.91 ; NH · -0.16, -0.66 ; NHCHO, 0.19, 0.00 ...
  49. [49]
    Hammett equation | Opinion - Chemistry World
    Jan 14, 2020 · Hammett was interested in this linear relationship between the rate constant of one reaction and the equilibrium constant of another.Missing: paper | Show results with:paper
  50. [50]
    Polar and Steric Substituent Constants for Aliphatic and o-Benzoate ...
    Density-based quantification of steric effects: validation by Taft steric parameters from acid-catalyzed hydrolysis of esters. Physical Chemistry Chemical ...