Fact-checked by Grok 2 weeks ago

Regioselectivity

Regioselectivity is a fundamental concept in that describes the preference of a to occur at one specific position or site within a over alternative positions, resulting in the predominant formation of one regioisomer among possible products. This selectivity arises when an unsymmetrical reagent, such as HCl, reacts with an unsymmetrical substrate like propene, yielding primarily 2-chloropropane rather than 1-chloropropane as the major product. The phenomenon is governed by principles like , which predicts that in electrophilic additions to alkenes, the attaches to the carbon with more hydrogens, favoring the more stable intermediate. In broader terms, regioselectivity is influenced by factors such as steric hindrance, electronic effects, and the three-dimensional structure of the molecule, making it essential for controlling reaction outcomes in . For instance, in Diels-Alder reactions involving 1-substituted dienes, the "ortho/para" rule directs the formation of specific cycloadducts, enhancing precision in constructing complex frameworks. Unlike , which concerns the spatial arrangement of atoms (e.g., vs. ), regioselectivity focuses on the constitutional differences in product , though both contribute to overall reaction efficiency. A reaction is termed regiospecific if it yields only one regioisomer exclusively, while regioselective reactions produce a major product alongside minor alternatives. The importance of regioselectivity extends to advanced applications in and , where achieving high regioregularity—complete preference for one positional —improves material properties like in conjugated polymers. In catalytic oxidations within confined environments like zeolites, regioselectivity can be amplified compared to homogeneous conditions, directing reactions to specific carbon positions for selective functionalization. Understanding and manipulating regioselectivity through design or reaction conditions remains a cornerstone of modern synthetic strategies, enabling the efficient assembly of pharmaceuticals and functional materials.

Fundamentals

Definition

Regioselectivity is a in describing the preferential formation of one regioisomer over others in a involving an unsymmetrical , where the occurs at one specific constitutional position rather than alternative possible sites. This selectivity arises when multiple orientations of bond formation or breaking are feasible, but predominates due to thermodynamic or kinetic factors, leading to a where the favored product constitutes a significant majority—potentially up to 100% in highly selective cases. The term encompasses a range of s, originally focused on additions to unsymmetrical alkenes but now broadly applied to various transformations in . Regioisomers are a subclass of constitutional isomers, characterized by having identical molecular formulas and overall connectivity but differing in the precise position of functional groups, double bonds, or other structural features within the carbon skeleton. For instance, in aromatic substitution, - and para-substituted products represent regioisomers of a monosubstituted derivative. This positional variation can significantly influence the compound's physical properties, reactivity, and , making regioselectivity a critical consideration in synthetic design to target specific isomers efficiently. A basic illustration of regioselectivity can be seen in the step of an unsymmetrical , such as propene (CH₃CH=CH₂). Here, the proton (H⁺) may add to either the terminal carbon (C1) or the substituted carbon (C2), potentially yielding a primary (CH₃CH₂CH₂⁺) or a secondary (CH₃CH⁺CH₃), respectively. Due to the inherent stability differences between these intermediates, the reaction favors protonation at C1, resulting in the secondary carbocation and thus one predominant regioisomeric pathway over the alternative. The term "regioselectivity" was coined in 1968 by Alfred Hassner to provide a precise for the effects observed in and elimination reactions, including cycloadditions and ring openings, distinguishing it from which pertains to spatial arrangements rather than positional preferences.

Relation to Other Selectivity Concepts

Regioselectivity is one of several key selectivity concepts in that guide reaction outcomes, but it specifically addresses the preferential formation of one constitutional (regioisomer) over another possible positional variants. This contrasts with , which favors one stereoisomer (such as diastereomers or ) based on spatial arrangement rather than connectivity. , on the other hand, involves the preferential reactivity of one over others in a multifunctional , without necessarily producing regioisomers. Enantioselectivity represents a of , emphasizing the asymmetric preference for one mirror-image in reactions involving chiral centers or environments. The distinctions among these concepts are summarized in the following table:
Selectivity TypePrimary FocusPreferred Products/Isomers
RegioselectivityPositional orientation of bond formation or breakingOne constitutional (regio)isomer over others
Preference among distinct sReaction at one functional group, yielding distinct products
Spatial (three-dimensional) arrangementOne stereoisomer (e.g., ) over another
EnantioselectivityChiral bias in stereoisomer formationOne over its mirror image
These selectivity types often coexist within the same reaction, where regioselectivity determines the site of reaction while governs the configuration at that site; for example, ring-opening reactions under basic conditions exhibit regioselectivity by favoring nucleophilic attack at the less substituted carbon and stereoselectivity through inversion of configuration at the attacked carbon. In organic synthesis, mastery of regioselectivity is essential for predicting product distributions and designing pathways that minimize isomer mixtures, thereby enabling precise control over molecular architecture and reducing ambiguity in identifying target compounds.

Key Principles

Markovnikov's Rule

Markovnikov's rule is a foundational principle in organic chemistry that predicts the regioselectivity of electrophilic addition reactions involving hydrogen halides (HX, where X is a halogen) and unsymmetrical alkenes. It states that in such additions, the hydrogen atom from HX adds to the carbon of the double bond that already has more hydrogen atoms attached, while the halogen atom adds to the carbon with fewer hydrogen atoms. This rule ensures the formation of the more stable alkyl halide product. The rule was first proposed by Russian chemist Vladimir Vasilyevich Markovnikov in his 1869 doctoral thesis at Kazan University, based on empirical observations of reactions between unsaturated hydrocarbons and hydrogen halides. The underlying mechanism of Markovnikov addition proceeds via a two-step process. First, the alkene's π-bond attacks the electrophilic hydrogen of HX, forming a intermediate at the more substituted carbon, which is more stable due to and inductive effects from adjacent alkyl groups (e.g., a secondary is preferred over a primary one). In the second step, the (X⁻) attacks the , yielding the product. For example, the addition of HX to propene (CH₃CH=CH₂) generates a secondary at the middle carbon, leading predominantly to 2-halopropane (CH₃CHXCH₃), rather than the primary carbocation-derived 1-halopropane (CH₃CH₂CH₂X). This pathway explains the rule's regioselectivity, as the leading to the more stable intermediate has lower energy. In practice, Markovnikov addition exhibits high regioselectivity, with the predicted product often comprising more than 95% of the mixture in favorable cases, such as when the stability difference is significant (e.g., secondary vs. primary). This selectivity can be inverted under free radical conditions, such as in the presence of peroxides for HBr additions, yielding the anti-Markovnikov product, though this mechanism differs fundamentally from the ionic pathway of the classical rule.

Other Classical Rules

Baldwin's rules, proposed in 1976, offer empirical guidelines for predicting the feasibility of ring closure reactions involving reactive intermediates like enolates, radicals, and carbocations. These rules categorize cyclizations based on the hybridization of the reacting center (trigonal, tetrahedral, or digonal) and the trajectory of the bond formation ( or relative to the ring). Favored processes minimize torsional in the ; for instance, 5--tet cyclizations are preferred over 5--tet due to reduced eclipsing interactions in the exo approach. In Baeyer-Villiger oxidations, regioselectivity follows the migratory aptitude of the substituents adjacent to the ketone carbonyl, with the empirical order tertiary alkyl > secondary alkyl ≈ aryl > primary alkyl > methyl, reflecting the group's capacity to bear positive charge during migration. For acetophenone, peracid oxidation yields phenyl acetate as the major product, with the aryl group migrating preferentially: \ce{C6H5C(O)CH3 + RCO3H -> C6H5OC(O)CH3 + RCO2H + H2O} This outcome aligns with the higher aptitude of phenyl over methyl. These classical rules function as practical, empirical tools for forecasting regioselectivity across diverse reaction types, from cyclizations to oxidations, though their application is context-specific.

Influencing Mechanisms

Electronic Factors

Electronic factors govern regioselectivity by influencing the distribution and the energetic preferences of transition states or intermediates through orbital overlap, charge stabilization, and effects. In pericyclic reactions like the Diels-Alder cycloaddition, explains regioselectivity via the interaction between the highest occupied () of the and the lowest unoccupied (LUMO) of the dienophile, where the largest orbital coefficients dictate the favored orientation for bond formation. This coefficient matching, as articulated by Fukui, leads to "ortho-like" or "para-like" adducts in unsymmetrical cases by maximizing stabilization from secondary orbital interactions. The relative stability of ionic intermediates significantly drives regioselectivity in reactions involving carbocations or carbanions. For carbocations, tertiary positions are preferred over primary due to enhanced stabilization from , where adjacent C-H σ bonds donate electron density to the empty p-orbital, and inductive effects from alkyl groups, though recent analyses show alkyl substituents are mildly electron-withdrawing inductively but overwhelmingly compensates for . This ordering—tertiary > secondary > primary—determines regioselective pathways in solvolysis or reactions. Similar electronic principles apply to anions, where electron-withdrawing groups stabilize negative charge through inductive withdrawal or delocalization. Polarization effects from substituents manifest prominently in , where directing groups alter ring via . directors like alkoxy groups increase negative charge density at and positions through donation, facilitating attack there, while meta directors such as nitro groups concentrate the highest negative charge at the meta position, as revealed by charge distribution analyses. This resonance-driven polarization ensures regioselective product formation without invoking simplistic donor-acceptor classifications. In allylic systems, charge delocalization via exemplifies electronic control over regioselectivity, as the positive charge in an allyl spreads across the π framework, rendering terminal carbons equivalent in reactivity. The key resonance structures are: \ce{CH2=CH-CH2^+ \leftrightarrow ^{+}CH2-CH=CH2} This delocalization, contributing approximately 20 kcal/mol to stability, influences regioselective allylic rearrangements or substitutions by favoring paths that maintain the resonant intermediate. Quantum chemical methods, particularly (DFT), offer insights into electronic factors by predicting regioselectivity ratios through energy calculations, often reproducing experimental selectivities for cycloadditions and substitutions with high accuracy. Such computations highlight how subtle differences in orbital energies or charge distributions dictate product ratios. This electronic preference is exemplified in applications like for protonation, where the more substituted forms preferentially.

Steric and Conformational Factors

Steric hindrance plays a pivotal role in regioselectivity by impeding the approach of to more congested sites in a , often directing reactions toward less substituted positions. In additions, such as the radical addition of thiols to 1-s (H₂C=CHR), the steric bulk of the R-substituent correlates with the Taft steric parameter (Eₛ), favoring anti-Markovnikov addition at the terminal carbon to minimize repulsion in the unsymmetrical . Similarly, in of s, the bulky reagent (e.g., 9-BBN) preferentially adds to the less hindered terminal carbon of terminal s, enhancing regioselectivity over electronic factors alone due to increased steric crowding. Conformational control further governs regioselectivity in cyclic systems, where preferred ring geometries dictate accessible transition states. For instance, in ring-opening reactions within polycyclic frameworks, chair-like conformations typically favor exo-selective openings, forming smaller rings via spiro transition states, as these minimize compared to boat-like states required for endo selectivity. In cyclohexene oxide, nucleophilic ring opening under basic conditions proceeds preferentially through a trans-diaxial pathway, where the axial approach aligns with the conformation to avoid excessive steric interactions in the . Steric effects can override electronic preferences, inverting expected regioselectivity and leading to high selectivity ratios. In the Mizoroki-Heck reaction involving methyl acrylate insertion into Pd-aryl bonds, bulky ligands such as 2,6-di(isopropyl)phenyl groups destabilize the electronically favored 2,1-insertion by ~9 kJ/mol, resulting in ~90% yield of the sterically preferred 1,2-insertion product. This inversion highlights how physical bulk can dominate over charge distribution, complementing electronic influences in overall selectivity. According to , regioselectivity arises from the lower energy pathway that minimizes steric repulsions, such as 1,3-diaxial interactions in cyclic . In aldol reactions following the Zimmerman-Traxler model, E-enolates favor anti-1,2 products because the positions substituents pseudo-equatorially, reducing 1,3-diaxial clashes between the R group and moieties compared to the syn pathway. These minimized interactions lower the , enhancing the kinetic preference for the observed regioisomer.

Reaction Examples

Addition Reactions

Addition reactions to unsaturated bonds exemplify regioselectivity, where the orientation of addition is dictated by electronic and steric factors, often following classical rules like Markovnikov's. In electrophilic additions, such as the acid-catalyzed hydration of alkenes, the proton from the acid adds to the less substituted carbon of the double bond, forming a carbocation at the more substituted position; subsequent nucleophilic attack by water yields the Markovnikov alcohol product, with the hydroxyl group on the more substituted carbon. For instance, hydration of propene primarily produces 2-propanol rather than 1-propanol, reflecting the stability of the secondary carbocation intermediate over the primary one. This process is highly regioselective, typically achieving greater than 95% yield of the Markovnikov product in simple cases due to the energetic preference for the more stable carbocation. Nucleophilic additions to conjugated systems, particularly α,β-unsaturated carbonyl compounds like enones, demonstrate competing regioselectivities between 1,2-addition at the carbonyl and 1,4-conjugate addition at the β-carbon. Under kinetic control, employing hard nucleophiles such as Grignard reagents at low temperatures, the 1,2-addition predominates because the direct attack on the electrophilic carbonyl is faster, often with selectivities of 80-95% favoring the allylic alcohol product. For example, methylmagnesium bromide adds to cyclohexenone primarily at the carbonyl carbon, yielding 1-methylcyclohex-2-en-1-ol as the kinetic product. In contrast, thermodynamic control, achieved with softer nucleophiles or under equilibrating conditions, shifts selectivity toward the 1,4-adduct, where the is protonated to restore the carbonyl, providing the β-substituted ketone. This duality arises from the resonance-stabilized intermediate, allowing reversion of the initial 1,2-adduct. Halogenation reactions in aqueous media, forming bromohydrins, also exhibit regioselectivity in unsymmetrical alkenes despite the anti addition . The electrophilic forms a bromonium ion intermediate, and then attacks the more substituted carbon, which bears greater positive charge character, placing the on the less substituted carbon. For propene, this yields 1-bromopropan-2-ol as the major product with near-complete regioselectivity (>98%), avoiding the alternative 2-bromopropan-1-ol. The selectivity stems from the partial -like during nucleophilic opening of the bromonium ion, aligning with electronic preferences similar to those in mechanisms.

Rearrangement and Cyclization Reactions

In the pinacol rearrangement, a vicinal diol undergoes acid-catalyzed dehydration to form a carbonyl compound, with regioselectivity governed by the migratory aptitude of the group antiperiplanar to the leaving water molecule in the intermediate carbocation. The order of migratory aptitude typically follows H > phenyl > tertiary alkyl > secondary alkyl > primary alkyl > methyl, leading to preferential migration of the more stable or electron-donating group, such as aryl over alkyl in unsymmetrical diols. For example, in the rearrangement of 1,2-diphenylethane-1,2-diol, a phenyl group migrates preferentially, yielding diphenylacetaldehyde in high yield under standard sulfuric acid conditions. This selectivity often exceeds 95% for the favored regioisomer when the diol is designed with clear aptitude differences, enabling predictable synthesis of aldehydes or ketones from complex polyols. The Diels-Alder cycloaddition exemplifies regioselectivity in pericyclic reactions, where unsymmetrical s and s combine to form cyclohexenes with specific substitution patterns dictated by frontier interactions. In reactions involving a 1-substituted and an electron-withdrawing-substituted , the "" orientation predominates, analogous to , resulting in the electron-donating and withdrawing groups being adjacent in the product. For instance, the reaction of 1-methoxybuta-1,3- with yields the 3-methoxycyclohexene-1-carbaldehyde ( product) as the major , often with >90% regioselectivity under thermal conditions without catalysts. The endo rule, favoring approach with maximum orbital overlap, further reinforces this regiochemical preference in concerted transition states, though it primarily influences . In the Baeyer-Villiger oxidation, ketones are converted to esters or lactones via peracid insertion of oxygen, with regioselectivity determined by the migratory aptitude of the adjacent alkyl or aryl groups in the Criegee intermediate. The general reaction can be represented as: \ce{R-C(=O)-R' ->[mCPBA] R-C(=O)-O-R'} where the group with higher migratory aptitude (typically tertiary > secondary ≈ aryl > primary > methyl) migrates, inserting oxygen adjacent to the less substituted or more electron-rich carbon. For cyclohexanone, symmetric migration yields caprolactone quantitatively, but in 2-methylcyclohexanone, the tertiary carbon migrates preferentially, affording the 7-methyl-substituted lactone in >95% regioselectivity using m-chloroperbenzoic acid. This control is crucial for synthesizing complex lactones in natural product synthesis, with aptitudes correlating to the ability to stabilize positive charge during migration. In cyclization reactions leading to ring formation, such as halocyclizations or metal-catalyzed variants, regioselectivity can also be influenced by , which favor exocyclic over endocyclic closure based on steric and torsional factors in five- or six-membered rings.

Modern Considerations

Synthetic Control Methods

In , controlling regioselectivity is essential for directing reactions toward desired products, and synthetic methods have evolved to manipulate this through targeted interventions. Catalyst design plays a pivotal role by enabling the tuning of electronic and steric properties to favor specific addition pathways. For instance, ligand-modified catalysts can invert traditional regioselectivity patterns, such as achieving anti-Markovnikov addition in hydrofunctionalization reactions. A notable example is the use of rhodium- complexes, which mediate the anti-Markovnikov hydrofunctionalization of terminal olefins by stabilizing key intermediates that direct the to the less substituted carbon. This approach has been particularly effective for unactivated alkenes, enabling the formation of heterocycles like tetrahydrofurans with >97% regioselectivity under mild conditions (25°C), demonstrating how porphyrin ligands enhance the catalyst's ability to override Markovnikov tendencies. Protecting groups provide another powerful strategy for enforcing regioselectivity in multi-step syntheses, particularly with multifunctional substrates like polyols or aromatic systems. By temporarily masking reactive sites, these groups block undesired pathways, allowing selective functionalization at unprotected positions. In , for example, a directing such as tetrafluoropyridyl on can influence the regiochemistry of , enabling substitution at remote positions that deviate from classical rules. This method is employed in multi-step syntheses through iterative protection-deprotection cycles to ensure precise control without altering the core . Reaction conditions, such as and choice, offer kinetic and thermodynamic levers to modulate regioselectivity without modifying the or catalyst. Lower favor kinetic products by limiting equilibration, while higher promote thermodynamic products through reversible pathways. A classic illustration is the addition of HBr to conjugated dienes like 1,3-butadiene, where low-temperature conditions (-80°C) yield predominantly the 1,2-addition product (kinetic control, ~80% selectivity), whereas elevated (40°C) shift to the 1,4-addition product (thermodynamic control, ~80% selectivity) due to . polarity further tunes this by stabilizing ionic intermediates; polar solvents enhance thermodynamic control by facilitating delocalization. Computational has emerged as a proactive tool for designing regioselective pathways, leveraging (DFT) and to forecast outcomes before experimentation. In the 2010s, semiempirical methods informed by DFT accurately predicted regioselectivity in reactions, achieving up to 95% success rates. approaches, trained on reaction databases, extended this to broader predictions, such as site-selectivity in various reactions including and cross-coupling, with top-3 accuracy of ~87% and top-5 of ~91%. These methods have streamlined the design of selective syntheses by integrating electronic factors into predictive algorithms.

Exceptions and Limitations

While classical rules like Markovnikov's predict regioselectivity based on stability in electrophilic s, exceptions arise in -mediated processes, such as the of to alkenes in the presence of peroxides, known as the Kharasch effect. This mechanism involves initiation by peroxide decomposition, leading to anti-Markovnikov orientation where adds to the more substituted carbon and to the less substituted one, as the preferentially abstracts from HBr due to bond strength considerations. Similarly, of alkenes with (BH3) proceeds via a concerted, syn that favors anti-Markovnikov regioselectivity, with attaching to the less substituted carbon due to steric and electronic factors in the four-center . Quantum mechanical effects can further deviate from predicted regioselectivity in certain reactions. For instance, quantum tunneling allows particles to penetrate energy barriers rather than surmount them, potentially favoring pathways to products disfavored under classical kinetics. Non-classical carbocations, such as cyclopropylcarbinyl ions, exhibit bridged structures that delocalize positive charge symmetrically across multiple carbons, affecting regioselectivity in ring-opening reactions and leading to predictable product outcomes from structural variations. Post-2010 research on directed C-H activation highlights ongoing variability in regioselectivity, particularly when weak coordinating groups are used to functionalize similar C-H bonds in complex arenes. For example, in palladium-catalyzed reactions of 1-substituted naphthalenes, directing groups can lead to mixtures at C-8 versus C-4 positions due to competing coordination modes, underscoring incomplete control in polysubstituted systems. This variability persists in aliphatic C(sp3)-H activations, where subtle conformational differences result in unpredictable despite directing strategies. Predicting regioselectivity becomes limited when electronic, steric, and factors compete, often yielding isomeric mixtures rather than high selectivity. In symmetric or near-symmetric substrates, such as disubstituted alkenes with balanced substituents, additions can produce 50:50 mixtures, as classical rules fail to differentiate pathways adequately. Computational models, while improving, still struggle with these cases due to the need for high-level quantum descriptors to capture dynamic effects, leaving gaps in forecasting outcomes for multifunctional molecules.

References

  1. [1]
    Regioselective - Chemistry LibreTexts
    Jan 22, 2023 · Regioselectiviy occurs in chemical reactions where one reaction site is preferred over another. For example, the addition of an asymmetric reagent (such as H- ...
  2. [2]
    Ch 6 : Regio- and Stereoselectivity - University of Calgary
    Regioselectivity. Addition of non-symmetrical reagent (i.e. A-B where A is not equal to B) to a non-symmetrical alkene (i.e. where the groups at each end of ...
  3. [3]
    Regioselectivity - an overview | ScienceDirect Topics
    Regioselectivity refers to the preference of a chemical reaction to occur at a specific position or region within a molecule, often influenced by the three- ...
  4. [4]
    regioselectivity (R05243) - IUPAC Gold Book
    (Originally the term was restricted to addition reactions of unsymmetrical reagents to unsymmetrical alkenes.) In the past, the term 'regiospecificity' was ...
  5. [5]
    Structural and vibrational spectroscopic analysis of anticancer drug ...
    Among the three categories of constitutional isomers, mitotane belongs to positional isomer which is also called regioisomer. ... position of the functional group ...
  6. [6]
    Illustrated Glossary of Organic Chemistry - Regioselective
    Regioselective: Any process that favors bond formation at a particular atom over other possible atoms. The description of a reaction's regioselectivity (or ...
  7. [7]
    Chemoselective or Regioselective? - Chemistry Europe - Wiley
    Mar 27, 2025 · Of the two concepts of selectivity, regioselectivity is the older. The terms regiospecific and regioselective were first proposed in 1968 by ...
  8. [8]
    Stereoselective and Stereospecific Reactions
    Apr 10, 2025 · A stereoselective reaction is selective for the formation of one stereoisomer over another. An enantioselective reaction is selective for the ...
  9. [9]
    15.8: Opening of Epoxides - Chemistry LibreTexts
    May 30, 2020 · The opening of expoxides is a regioselective reaction depending on the structure of the epoxide and reaction conditions. The stereochemistry ...
  10. [10]
    Computational tools for the prediction of site- and regioselectivity of ...
    Mar 4, 2025 · The regio- and site-selectivity of organic reactions is one of the most important aspects when it comes to synthesis planning.
  11. [11]
    [PDF] Towards the 150th Anniversary of the Markovnikov Rule - AMyD
    In his doctoral thesis, in 1869, Vladimir Markovnikov. (Figure 1) formulated the famous rule that appears in almost every textbook on organic chemistry—be it ...
  12. [12]
  13. [13]
    7.8 Orientation of Electrophilic Additions: Markovnikov's Rule |
    Sep 20, 2023 · ... Vladimir Markovnikov proposed in 1869 what has become known as: Markovnikov's rule. In the addition of HX to an alkene, the H attaches to the ...Missing: original | Show results with:original
  14. [14]
  15. [15]
    Rules for ring closure - Journal of the Chemical ... - RSC Publishing
    Three rules which have been found useful, on an empirical basis, to predict the relative facility of ring forming reactions are presented.<|separator|>
  16. [16]
    [PDF] Kenichi Fukui - Nobel Lecture
    Fukui. 13. By considering the HOMO-LUMO interactions between the fragments of a conjugated chain divided into parts,21 the frontier orbital theory can yield.Missing: regioselectivity | Show results with:regioselectivity
  17. [17]
    Stability of alkyl carbocations - RSC Publishing
    Oct 6, 2022 · We reveal the exact source of the stability trend of carbocations by using partial reactions in a thermodynamic cycle quantitatively decomposing the effect of ...
  18. [18]
    Hyperconjugation: A More Coherent Approach - ACS Publications
    May 15, 2012 · This article is a discussion of hyperconjugation and a proposal for introducing the idea within the framework of other important concepts such as conjugation ...
  19. [19]
    Time to stop mentioning alkyl group inductive effects - Chemistry World
    Jan 6, 2025 · New computational evidence reiterates that alkyl groups are inductively electron-withdrawing and that hyperconjugation dominates.
  20. [20]
    Where does the electron go? The nature of ortho/para and meta ...
    Nov 21, 2014 · Ortho/para directing groups have most negative charges on the ortho/para positions, while meta directing groups often possess the largest negative charge on ...
  21. [21]
    Resonance Energies of the Allyl Cation and Allyl Anion: Contribution ...
    This is because a full negative charge is delocalized in the allyl anion, between two equivalent resonance structures; resonance in the neutral acid, on the ...Abstract · Computational Methods · Supporting Information...
  22. [22]
    Is DFT Accurate Enough to Calculate Regioselectivity? The Case of ...
    Mar 10, 2023 · We investigated whether DFT calculations are an accurate tool to predict the regioselectivity of uncatalyzed thermal azide 1,3-DCs.
  23. [23]
  24. [24]
    The transition metal-catalysed hydroboration reaction
    Oct 7, 2022 · This regioselectivity is often even higher than with the monosubstituted aliphatic alkenes due to increased steric crowding, which also results ...
  25. [25]
  26. [26]
    Breaking the regioselectivity rule for acrylate insertion in the ... - PNAS
    May 11, 2011 · The regioselectivity of insertion follows the same pattern for both reactions. Electron-deficient olefins [e.g., methyl acrylate (MA)] ...Abstract · Results And Discussion · Materials And Methods
  27. [27]
  28. [28]
  29. [29]
    Addition Reactions of Alkenes - MSU chemistry
    Regioselectivity and the Markovnikov Rule ... In both cases the overall reaction is the addition of water to the double bond, but the regioselectivity is reversed ...
  30. [30]
  31. [31]
    1,2- vs 1,4-Regioselectivity of Lithiated Phenylacetonitrile toward α,β ...
    In THF, the monomeric bridged ion pair 2 leads to predominately 1,2-addition with α,β-unsaturated carbonyl compounds. High 1,4-regioselectivity is evidenced ...Missing: alpha | Show results with:alpha
  32. [32]
    Reaction of Alkenes with Bromine - Chemistry Steps
    When an unsymmetrical alkene is used, the nucleophilic attack by water occurs at the more substituted carbon of the bromonium ion, and the halogen ends up on ...
  33. [33]
    Theoretical study of the reaction mechanism and migratory aptitude ...
    Theoretical study of the reaction mechanism and migratory aptitude of the pinacol rearrangement ... Regioselective 1,2-Diol Rearrangement by Controlling ...Missing: seminal | Show results with:seminal
  34. [34]
    THE PINACOL-PINACOLIN REARRANGEMENT. THE RELATIVE ...
    THE PINACOL-PINACOLIN REARRANGEMENT. THE RELATIVE MIGRATION APTITUDES OF ARYL GROUPS1.Missing: regioselectivity seminal
  35. [35]
    the case of the ortho–para regioselectivity rule in Diels–Alder reactions
    Jan 9, 2020 · The regioselectivity of the Diels–Alder reaction is predicted by the ortho–para rule which has been explained from FMO theory.Missing: cycloaddition meta seminal
  36. [36]
    Anti‐Markovnikov Hydrofunctionalization of Olefins Mediated by ...
    Jan 21, 2004 · Porphyrin–rhodium complexes have been shown to exhibit remarkable reactivity and selectivity for each step of the proposed catalytic cycle (see ...
  37. [37]
    Protecting Group-Controlled Remote Regioselective Electrophilic ...
    May 7, 2020 · Being able to utilize a protecting group to influence remote regiocontrol offers a simple alternative approach to direct late-stage ...
  38. [38]
    Reactions of Dienes: 1,2 and 1,4 Addition - Master Organic Chemistry
    Mar 22, 2017 · Kinetic and thermodynamic control in addition of HBr to dienes can give 1,2- and 1,4- products. How do we explain this?
  39. [39]
    [PDF] Guiding Chemical Synthesis: Computational Prediction of the ... - arXiv
    The prediction of the regioselectivity of CH functionalization is usually done using heuristic rules that can be found in most organic chemistry textbooks, such ...
  40. [40]
    Prediction of Organic Reaction Outcomes Using Machine Learning
    Apr 18, 2017 · A combination of traditional reaction templates and machine learning enables the prediction of organic reaction products in silico using open source data from ...
  41. [41]
    Tunneling Control of Chemical Reactions: The Third Reactivity ...
    The term denotes a tunneling reaction that passes through a high but narrow potential energy barrier, leading to formation of a product that would be disfavored ...
  42. [42]
    Taming nonclassical carbocations to control small ring reactivity
    Jan 12, 2024 · Here, we show that such reaction outcomes can be controlled by subtle changes to the structure of nonclassical carbocation.
  43. [43]
    Transition-Metal-Catalyzed C–H Bond Activation for the Formation of ...
    This minireview summarizes reactions that allow regioselective functionalization of 1-substituted naphthalenes based on directed C-H activation strategies.Missing: variability | Show results with:variability
  44. [44]
    Achieving Site-Selectivity for C–H Activation Processes Based on ...
    Abstract. The ability to differentiate between highly similar C–H bonds in a given molecule remains a fundamental challenge in organic chemistry.Missing: post- | Show results with:post-
  45. [45]
    Computational tools for the prediction of site- and regioselectivity of ...
    Mar 4, 2025 · Abstract. The regio- and site-selectivity of organic reactions is one of the most important aspects when it comes to synthesis planning.
  46. [46]
    No electron left behind: a rule-based expert system to predict ... - NIH
    Among the most fundamental problems in organic chemistry is predicting the course and major products of arbitrary reactions. In addition to being a ...