Stereoelectronic effect

Last updated

In chemistry, primarily organic and computational chemistry, a stereoelectronic effect [1] is an effect on molecular geometry, reactivity, or physical properties due to spatial relationships in the molecules' electronic structure, in particular the interaction between atomic and/or molecular orbitals. [2] Phrased differently, stereoelectronic effects can also be defined as the geometric constraints placed on the ground and/or transition states of molecules that arise from considerations of orbital overlap. [3] Thus, a stereoelectronic effect explains a particular molecular property or reactivity by invoking stabilizing or destabilizing interactions that depend on the relative orientations of electrons (bonding or non-bonding) in space. [4]

Contents

Stereoelectronic effects present themselves in other well-known interactions. These include important phenomena such as the anomeric effect and hyperconjugation. It is important to note that stereoelectronic effects should not be misunderstood as a simple combination of steric effects and electronic effects.

Founded on a few general principles that govern how orbitals interact, the stereoelectronic effect, along with the steric effect, inductive effect, solvent effect, mesomeric effect, and aromaticity, is an important type of explanation for observed patterns of selectivity, reactivity, and stability in organic chemistry. In spite of the relatively straightforward premises, stereoelectronic effects often provide explanations for counterintuitive or surprising observations. As a result, stereoelectronic factors are now commonly considered and exploited in the development of new organic methodology and in the synthesis of complex targets. The scrutiny of stereoelectronic effects has also entered the realms of biochemistry and pharmaceutical chemistry in recent years.

A stereoelectronic effect generally involves a stabilizing donor-acceptor (i.e., filled bonding-empty antibonding, 2-electron 2-orbital) interaction. The donor is usually a higher bonding or nonbonding orbital and the acceptor is often a low-lying antibonding orbital as shown in the scheme below. Whenever possible, if this stereoelectronic effect is to be favored, the donor-acceptor orbitals should have (1) a small energy gap and (2) be geometrically well disposed for interaction. In particular, this means that the shapes of the donor and acceptor orbitals (including π or σ symmetry and size of the interacting lobes) must be well-matched for interaction; an antiperiplanar orientation is especially favorable. Some authors require stereoelectronic effects to be stabilizing. [1] However, destabilizing donor-donor (i.e., filled bonding-filled antibonding, 4-electron 2-orbital) interactions are occasionally invoked and are also sometimes referred to as stereoelectronic effects, although such effects are difficult to distinguish from generic steric repulsion. [3] [5]

Tim Project Image 1.gif

Trend of different orbitals

Take the simplest CH2X–CH3 system as an example; the donor orbital is σ(C–H) orbital and the acceptor is σ*(C–X). When moving from fluorine to chlorine, then to bromine, the electronegativity of the halogen and the energy level of the σ*(C–X) orbitals decreases. [6] Consequently, the general trend of acceptors can be summarized as: π*(C=O)>σ*(C–Hal)>σ*(C–O)>σ*(C–N)>σ*(C–C), σ*(C–H). For donating orbitals, the nonbonding orbitals, or the lone pairs, are generally more effective than bonding orbitals due to the high energy levels. Also, different from acceptors, donor orbitals require less polarized bonds. Thus, the general trends for donor orbitals would be: n(N)>n(O)>σ(C–C), σ(C–H)>σ(C–N)>σ(C–O)>σ(C–S)>σ(C–Hal). [5]

Tim Project Image 2.gif

Stereoelectronic effect can be directional in specific cases. The radius of sulfur is much larger than the radius of carbon and oxygen. Thus the differences in C–S bond distances generate a much-amplified difference in the two stereoelectronic effects in 1,3-dithiane (σ(C–H) → σ*(C–S)) than in 1,3-dioxane(σ(C–H) → σ*(C–O)). [6] The differences between C–C and C–S bonds shown below causes a significant difference in the distances between C–S and two C–H bonds. The shorter the difference is, the better the interaction and the stronger the stereoelectronic effect. [6]

Tim Project Image 3.gif

Influence on stability

If there is an electropositive substituent (e.g. –SiR3, –SnR3, –HgR, etc.) at the β-position of carbocation, the positive charge could be stabilized which is also due largely to the stereoelectronic effect (illustrated below using –SiR3 as an example). The orientation of the two interacting orbitals can have a significant effect on the stabilization effect (σ(C–Si) → empty p orbital), where antiperiplanar (180°) > perpendicular (90°) > syn (0°). [7]

Tim Project Image 4.gif

Influence on conformation

Gauche effect

One structural consequence of acyclic systems due to the stereoelectronic effect is the gauche effect. [8] In 1,2-difluoroethane, despite the steric clash, the preferred conformation is the gauche one because σ(C–H) is a good donor and σ*(C–F) is a good acceptor and the stereoelectronic effect (σ(C–H) → σ*(C–F)) requires the energy minimum to be gauche instead of anti. [9]

Tim Project Image 5.gif

This gauche effect and its impact on conformation are important in biochemistry. For example, in HIF-α subunit fragments containing (2S,4R)-4-hydroxyproline, the gauche interaction favors the conformer that can bind to the active site of pVHL. [10] pVHL mediates the proteasomal degradation of HIF1A and with that the physiological response to hypoxia.

Tim Project Image 6.gif

Special effects of fluorine substituent

Stereoelectronic effects can have a significant influence in pharmaceutical research. Generally, the substitution of hydrogen by fluorine could be regarded as a way to tune both the hydrophobicity and the metabolic stability of a drug candidate. Moreover, it can have a profound influence on conformations, often due to stereoelectronic effects, in addition to normal steric effects resulting from the larger size of the fluorine atom. For instance, the ground state geometries of anisole (methoxybenzene) and (trifluoromethoxy)benzene differ dramatically. In anisole, the methyl group prefers to be coplanar with the phenyl group, while (trifluoromethoxy)benzene favors a geometry in which the [C(aryl)–C(aryl)–O–C(F3)] dihedral angle is around 90°. In other words, the O–CF3 bond is perpendicular to the plane of the phenyl group. [11]

Li Ti Dian Zi Xiao Ying Shu Tu -ZQK.png

Further studies illustrate that even for only one or two hydrogen atoms in a methyl group being replaced by a fluorine atom, the distortion in the structure can also be significant, with the [C(aryl)–C(aryl)–O–C(H2F)] dihedral angle in the energy minimized structure being around 24° and the [C(aryl)–C(aryl)–O–C(HF2)] dihedral angle 33°. [11]

Influence on reaction selectivity

Reductive cyclizations

Although the energy difference between coplanar anisole and its isomer is quite large, the rotation between the O–CH3 bond becomes favorable when the electronic properties of methoxy group on aromatic rings need to be altered to stabilize an unusual intermediate or a transition state. In the following reaction, the regioselectivity could be rationalized as the out-of-plane rotation of the O–C bond which changes the methoxy group from an in-plane donor group to an out-of-plane acceptor group. [12]

Tim Project Image 8.gif

The intermediate of the above reaction is the di-anion and the stereoelectronic effect that stabilizes this intermediate over the other one is the fact that the anionic charge at the para position could delocalize to the oxygen atom via orbital interaction: π(benzene) → σ*(O–CH3). [12]

Tim Project Image 9.gif

Hydrogenation

Even remote substituents on the benzene ring can affect the electron density on the aromatic ring and in turn influence the selectivity. In the hydrogenation of ketones using CBS catalysts, the ketone coordinates to the boron atom with the lone pair on the oxygen atom. In the following example, the inductive influence of the substituents can lead to differentiation of the two sp2 lone pairs on the oxygen atom. [13]

The relevant stereoelectronic interaction in the starting material is the nO → σ*(Ccarbonyl–Caryl) interaction. The electron-withdrawing substituent on the benzene ring depletes the electron density on the aromatic ring and thus makes the σ*(Ccarbonyl–Caryl(nitro)) orbital a better acceptor than σ*(Ccarbonyl–Caryl(methoxy)). These two stereoelectronic interactions use different lone pairs on the oxygen atom (the one antiperiplanar to the σ* in question for each), leading to lone pairs with different electron densities. In particular, the enhanced depletion of electron density from the lone pair antiperiplanar to the 4-nitrophenyl group leads to weakened ability for that lone pair to coordinate to boron. This in turn results in the lone pair antiperiplanar to the 4-methoxyphenyl binding preferentially to the catalyst, leading to well-defined facial selectivity. Under optimized conditions, the product is formed with excellent levels of enantioselectivity (95% ee). [13]

CBS-stereoelectronics.png

Influence on thermodynamics

Influence on equilibrium

The stereoelectronic effect influences the thermodynamics of equilibrium. For example, the following equilibrium could be achieved via a cascade of pericyclic reactions.

Tim Project Image 10.gif

Despite very similar structures, one of the two isomers is strongly favored over the other because of a stereoelectronic effect. Since the σ*C-C orbital adjacent to the electron-withdrawing carbonyl group is lower in energy and is therefore a better acceptor than the σ*C-C orbital adjacent to the methoxy, the isomer in which the nO(σ) lone pair is able to donate into this lower-energy antibonding orbital will be stabilized (orbital interaction illustrated). [14]

Another example of the preference in the equilibrium within the area of pericyclic reaction is shown below. The stereoelectronic effect that affect the equilibrium is the interaction between the delocalized “banana bonds” and the empty p orbital on the boron atom. [15]

Tim Project Image-Add-3.gif

Influence on resonance structures

In another case, the stereoelectronic effect can result in an increased contribution of one resonance structure over another, which leads to further consequences in reactivity. For 1,4-benzoquinone monoxime, there are significant differences in the physical properties and reactivities between C2-C3 double bond and C5-C6 double bond. For instance, in the 1H NMR, 3J23 higher than 3J56. [16] The C2-C3 double bond also selectively undergoes Diels–Alder reaction with cyclopentadiene, despite the increased steric hindrance on that side of the molecule. [17] These data illustrate an increased contribution of resonance structure B over structure A. The authors argue that the donation from nN to σ*C4-C3 orbital lengthens the C4–C3 bond (C4 is the carbon bearing the nitrogen substituent), which reduces the p-p overlap between these two atoms. This in turn decreasing the relative importance of structure A which has a double bond between C4 and C3. [18]

Tim Project Image 11.gif

Application in asymmetric Diels–Alder reactions

In the asymmetric Diels–Alder reactions, instead of using chiral ligands or chiral auxiliaries to differentiate the side selectivity of the dienolphiles, the differentiation of face selectivity of the dienes (especially for cyclopentadiene derivatives) using stereoelectronic effects have been reported by Woodward since 1955. [19] A systematic research of facial selectivity using substituted cyclopentadiene or permethylcyclopentadiene derivatives have been conducted and the results can be listed as below. [20]

Tim Project Image-Add-4.gif

The stereoelectronic effect affecting the outcome of the facial selectivity of the diene in the Diels–Alder reaction is the interaction between the σ(C(sp2)–CH3) (when σ(C(sp2)–X) is a better acceptor than a donor) or σ(C(sp2)–X) (when σ(C(sp2)–X) is a better donor than an acceptor) and the σ* orbital of the forming bond between the diene and the dienophile. [20]

Tim Project Image 16.gif

If the two geminal substituents are both aromatic rings with different substituents tuning the electron density, the differentiation of the facial selectivity is also facile where the dienophile approaches the diene anti to the more electron-rich C–C bond where the stereoelectronic effect, in this case, is similar to the previous one. [21]

Tim Project Image 17.gif

The ring opening of cyclobutene under heating conditions can have two products: inward and outward rotation.

Tim Project Image 18.gif

The inward rotation transition state of the structure shown below is relatively favored for acceptor R substituents (e.g. NO2) but is especially disfavored by donor R substituents (e.g. NMe2). [22]

Tim Project Image 19.gif

Stereoelectronic effect versus steric clash

Sometimes, stereoelectronic effects can win over extreme steric clash. In a similar cyclobutene ring-opening reaction, the trimethylsilyl group, which is very bulky, still favors the inward rotation. The stereoelectronic effect, which is the interaction shown above when the acceptor orbital is the σ*(Si–CH3), appears to be a more predominant factor in determining the reaction selectivity against the steric hindrance and even wins over the penalty of the disrupted conjugation system of the product due to steric clash. [23]

Furthermore, the acceptor orbitals are not limited to the antibonding orbitals of carbon-heteroatom bonds or the empty orbitals; in the following case, the acceptor orbital is the σ*(B–O) orbital. In the six-membered ring transition state, the stereoelectronic interaction is σ(C–X) → σ*(B–O). [24]

Tim Project Image 20.gif

Stereoelectronic effects in biomolecules

Molecular recognition events mediated through orbital interactions are critical in a number of biological processes such as enzyme catalysis. [25] Stabilizing interactions between proteins and carbohydrates in glycosylated proteins also exemplify the role of stereoelectronic effects in biomolecules. [26]

Related Research Articles

Pi backbonding Movement of electrons from one atoms orbital to a symmetric antibonding orbital on another

In chemistry, π backbonding, also called π backdonation, is when electrons move from an atomic orbital on one atom to an appropriate symmetry antibonding orbital on a π-acceptor ligand. It is especially common in the organometallic chemistry of transition metals with multi-atomic ligands such as carbon monoxide, ethylene or the nitrosonium cation. Electrons from the metal are used to bond to the ligand, in the process relieving the metal of excess negative charge. Compounds where π backbonding occurs include Ni(CO)4 and Zeise's salt. IUPAC offers the following definition for backbonding:

A description of the bonding of π-conjugated ligands to a transition metal which involves a synergic process with donation of electrons from the filled π-orbital or lone electron pair orbital of the ligand into an empty orbital of the metal (donor–acceptor bond), together with release (back donation) of electrons from an nd orbital of the metal (which is of π-symmetry with respect to the metal–ligand axis) into the empty π*-antibonding orbital of the ligand.

A transition metal carbene complex is an organometallic compound featuring a divalent organic ligand. The divalent organic ligand coordinated to the metal center is called a carbene. Carbene complexes for almost all transition metals have been reported. Many methods for synthesizing them and reactions utilizing them have been reported. The term carbene ligand is a formalism since many are not derived from carbenes and almost none exhibit the reactivity characteristic of carbenes. Described often as M=CR2, they represent a class of organic ligands intermediate between alkyls (−CR3) and carbynes (≡CR). They feature in some catalytic reactions, especially alkene metathesis, and are of value in the preparation of some fine chemicals.

Baeyer–Villiger oxidation

The Baeyer–Villiger oxidation is an organic reaction that forms an ester from a ketone or a lactone from a cyclic ketone, using peroxyacids or peroxides as the oxidant. The reaction is named after Adolf von Baeyer and Victor Villiger who first reported the reaction in 1899.

Hyperconjugation Concept in organic chemistry

In organic chemistry, hyperconjugation refers to the delocalization of electrons with the participation of bonds of primarily σ-character. Usually, hyperconjugation involves the interaction of the electrons in a sigma (σ) orbital with an adjacent unpopulated non-bonding p or antibonding σ* or π* orbitals to give a pair of extended molecular orbitals. However, sometimes, low-lying antibonding σ* orbitals may also interact with filled orbitals of lone pair character (n) in what is termed negative hyperconjugation. Increased electron delocalization associated with hyperconjugation increases the stability of the system. In particular, the new orbital with bonding character is stabilized, resulting in an overall stabilization of the molecule. Only electrons in bonds that are in the β position can have this sort of direct stabilizing effect — donating from a sigma bond on an atom to an orbital in another atom directly attached to it. However, extended versions of hyperconjugation can be important as well. The Baker–Nathan effect, sometimes used synonymously for hyperconjugation, is a specific application of it to certain chemical reactions or types of structures.

Quintuple bond

A quintuple bond in chemistry is an unusual type of chemical bond, first reported in 2005 for a dichromium compound. Single bonds, double bonds, and triple bonds are commonplace in chemistry. Quadruple bonds are rarer but are currently known only among the transition metals, especially for Cr, Mo, W, and Re, e.g. [Mo2Cl8]4− and [Re2Cl8]2−. In a quintuple bond, ten electrons participate in bonding between the two metal centers, allocated as σ2π4δ4.

The 3-center 4-electron (3c–4e) bond is a model used to explain bonding in certain hypervalent molecules such as tetratomic and hexatomic interhalogen compounds, sulfur tetrafluoride, the xenon fluorides, and the bifluoride ion. It is also known as the Pimentel–Rundle three-center model after the work published by George C. Pimentel in 1951, which built on concepts developed earlier by Robert E. Rundle for electron-deficient bonding. An extended version of this model is used to describe the whole class of hypervalent molecules such as phosphorus pentafluoride and sulfur hexafluoride as well as multi-center π-bonding such as ozone and sulfur trioxide.

Anomeric effect

In organic chemistry, the anomeric effect or Edward-Lemieux effect is a stereoelectronic effect that describes the tendency of heteroatomic substituents adjacent to a heteroatom within a cyclohexane ring to prefer the axial orientation instead of the less hindered equatorial orientation that would be expected from steric considerations. This effect was originally observed in pyranose rings by J. T. Edward in 1955 when studying carbohydrate chemistry.

Gauche effect

In the study of conformational isomerism, the Gauche effect is an atypical situation where a gauche conformation is more stable than the anti conformation (180°).

Asymmetric induction

In stereochemistry, asymmetric induction describes the preferential formation in a chemical reaction of one enantiomer or diastereoisomer over the other as a result of the influence of a chiral feature present in the substrate, reagent, catalyst or environment. Asymmetric induction is a key element in asymmetric synthesis.

The Hammett equation in organic chemistry describes a linear free-energy relationship relating reaction rates and equilibrium constants for many reactions involving benzoic acid derivatives with meta- and para-substituents to each other with just two parameters: a substituent constant and a reaction constant. This equation was developed and published by Louis Plack Hammett in 1937 as a follow-up to qualitative observations in a 1935 publication.

Allylic strain Type of strain energy in organic chemistry

Allylic strain in organic chemistry is a type of strain energy resulting from the interaction between a substituent on one end of an olefin with an allylic substituent on the other end. If the substituents are large enough in size, they can sterically interfere with each other such that one conformer is greatly favored over the other. Allyic strain was first recognized in the literature in 1965 by Johnson and Malhotra. The authors were investigating cyclohexane conformations including endocyclic and exocylic double bonds when they noticed certain conformations were disfavored due to the geometry constraints caused by the double bond. Organic chemists capitalize on the rigidity resulting from allylic strain for use in asymmetric reactions.

Carbon–fluorine bond Covalent bond between carbon and fluorine atoms

The carbon–fluorine bond is a polar covalent bond between carbon and fluorine that is a component of all organofluorine compounds. It is one of the strongest single bonds in chemistry, and relatively short, due to its partial ionic character. The bond also strengthens and shortens as more fluorines are added to the same carbon on a chemical compound. As such, fluoroalkanes like tetrafluoromethane are some of the most unreactive organic compounds.

An electronic effect influences the structure, reactivity, or properties of molecule but is neither a traditional bond nor a steric effect. In organic chemistry, the term stereoelectronic effect is also used to emphasize the relation between the electronic structure and the geometry (stereochemistry) of a molecule.

Flippin–Lodge angle

The Flippin–Lodge angle is one of two angles used by organic and biological chemists studying the relationship between a molecule's chemical structure and ways that it reacts, for reactions involving "attack" of an electron-rich reacting species, the nucleophile, on an electron-poor reacting species, the electrophile. Specifically, the angles—the Bürgi–Dunitz, , and the Flippin–Lodge, —describe the "trajectory" or "angle of attack" of the nucleophile as it approaches the electrophile, in particular when the latter is planar in shape. This is called a nucleophilic addition reaction and it plays a central role in the biological chemistry taking place in many biosyntheses in nature, and is a central "tool" in the reaction toolkit of modern organic chemistry, e.g., to construct new molecules such as pharmaceuticals. Theory and use of these angles falls into the areas of synthetic and physical organic chemistry, which deals with chemical structure and reaction mechanism, and within a sub-specialty called structure correlation.

Carbohydrate conformation refers to the overall three-dimensional structure adopted by a carbohydrate (saccharide) molecule as a result of the through-bond and through-space physical forces it experiences arising from its molecular structure. The physical forces that dictate the three-dimensional shapes of all molecules—here, of all monosaccharide, oligosaccharide, and polysaccharide molecules—are sometimes summarily captured by such terms as "steric interactions" and "stereoelectronic effects".

Oxocarbenium

An oxocarbeniumion is a chemical species characterized by a central sp2-hybridized carbon, an oxygen substituent, and an overall positive charge that is delocalized between the central carbon and oxygen atoms. A oxocarbenium ion is represented by two limiting resonance structures, one in the form of a carbenium ion with the positive charge on carbon and the other in the form of an oxonium species with the formal charge on oxygen. As a resonance hybrid, the true structure falls between the two. Compared to neutral carbonyl compounds like ketones or esters, the carbenium ion form is a larger contributor to the structure. They are common reactive intermediates in the hydrolysis of glycosidic bonds, and are a commonly used strategy for chemical glycosylation. These ions have since been proposed as reactive intermediates in a wide range of chemical transformations, and have been utilized in the total synthesis of several natural products. In addition, they commonly appear in mechanisms of enzyme-catalyzed biosynthesis and hydrolysis of carbohydrates in nature. Anthocyanins are natural flavylium dyes, which are stabilized oxocarbenium compounds. Anthocyanins are responsible for the colors of a wide variety of common flowers such as pansies and edible plants such as eggplant and blueberry.

In organic chemistry, the Cieplak effect is a predictive model to rationalize why nucleophiles preferentially add to one face of a carbonyl over another. Proposed by Andrzej Stanislaw Cieplak in 1980, it predicts anomalous results that other models of the time, such as the Cram and Felkin–Anh models, cannot justify. In the Cieplak model, electrons from a neighboring bond delocalize into the forming carbon–nucleophile (C–Nuc) bond, lowering the energy of the transition state and accelerating the rate of reaction. Whichever bond can best donate its electrons into the C–Nuc bond determines which face of the carbonyl the nucleophile will add to. The nucleophile may be a number of reagents, most commonly organometallic or reducing agents. The Cieplak effect is subtle, and often competes with sterics, solvent effects, counterion complexation of the carbonyl oxygen, and other effects to determine product distribution. Subsequent work has questioned its legitimacy.

Anti-periplanar Conformation of an A−B−C−D molecule where the A−B C−D dihedral angle is of magnitude >150°

In organic chemistry, anti-periplanar, or antiperiplanar, describes the A−B−C−D bond angle in a molecule. In this conformer, the dihedral angle of the A−B bond and the C−D bond is greater than +150° or less than −150°. Anti-periplanar is often used in textbooks to mean strictly anti-coplanar, with an A−BC−D dihedral angle of 180°. In a Newman projection, the molecule will be in a staggered arrangement with the anti-periplanar functional groups pointing up and down, 180° away from each other. Figure 5 shows 2-chloro-2,3-dimethylbutane in a sawhorse projection with chlorine and a hydrogen anti-periplanar to each other.

Digermyne

Digermynes are a class of compounds that are regarded as the heavier digermanium analogues of alkynes. The parent member of this entire class is HGeGeH, which has only been characterized computationally, but has revealed key features of the whole class. Because of the large interatomic repulsion between two Ge atoms, only kinetically stabilized digermyne molecules can be synthesized and characterized by utilizing bulky protecting groups and appropriate synthetic methods, for example, reductive coupling of germanium(II) halides.

Plumbylene

Plumbylenes (or plumbylidenes) are divalent organolead(II) analogues of carbenes, with the general chemical formula, R2Pb, where R denotes a substituent. Plumbylenes possess 6 electrons in their valence shell, and are considered open shell species.

References

  1. 1 2 Alabugin, I. V. Stereoelectronic Effects: the Bridge between Structure and Reactivity. John Wiley & Sons Ltd, Chichester, UK, 2016. http://eu.wiley.com/WileyCDA/WileyTitle/productCd-1118906349.html
  2. Cramer, Christopher J. (1996). "Hyperconjugation as it affects conformational analysis". Journal of Molecular Structure: THEOCHEM. 370 (2–3): 135–146. doi:10.1016/S0166-1280(96)04567-8. ISSN   0166-1280.
  3. 1 2 Evans, D. A. (2006). Chemistry 206 Lecture Notes. Cambridge, MA: Harvard University (Unpublished). pp. 1–2 (Lecture 1).
  4. Pierre., Deslongchamps (1983). Stereoelectronic effects in organic chemistry (1st ed.). Oxford [Oxfordshire]: Pergamon Press. ISBN   0080261841. OCLC   9412829.
  5. 1 2 Kirby, A. J. (1996). Stereoelectronic Effects. Oxford: Oxford University Press. ISBN   9780198558934.
  6. 1 2 3 Alabugin, Igor V.; Zeidan, Tarek A. (2002). "Stereoelectronic Effects and General Trends in Hyperconjugative Acceptor Ability of σ Bonds". Journal of the American Chemical Society. 124 (12): 3175–3185. doi:10.1021/ja012633z. ISSN   0002-7863. PMID   11902907.
  7. Carey, F. A.; Sundberg, R. J. (2007). Advanced Organic Chemistry, Part A: Structure and Mechanism (5th ed.). New York: Springer. ISBN   978-0387448978.
  8. Hanack, M. (1965). Conformational Theory. New York and London: Academic Press.
  9. Anslyn, E. V.; Dougherty, D. A. (2004). Modern Physical Organic Chemistry. Sausalito, Calif.: University Science Books. ISBN   1891389319.
  10. Fujimori, D.G. (2009). "Hypoxia sensing goes gauche". Nat. Chem. Biol. 5 (4): 202–203. doi:10.1038/nchembio0409-202. PMID   19295524.
  11. 1 2 Müller, K.; Faeh, C.; Diederich, F. (2007). "Fluorine in Pharmaceuticals: Looking Beyond Intuition". Science. 317 (5846): 1881–1886. Bibcode:2007Sci...317.1881M. doi:10.1126/science.1131943. ISSN   0036-8075. PMID   17901324.
  12. 1 2 Peterson, Paul W.; Shevchenko, Nikolay; Alabugin, Igor V. (2013). ""Stereoelectronic Umpolung": Converting a p-Donor into a σ-Acceptor via Electron Injection and a Conformational Change". Organic Letters. 15 (9): 2238–2241. doi:10.1021/ol400813d. ISSN   1523-7060. PMID   23639080.
  13. 1 2 Corey, E.J.; Helal, Christopher J. (1995). "Novel electronic effects of remote substituents on the oxazaborolidine-catalyzed enantioselective reduction of ketones". Tetrahedron Letters. 36 (50): 9153–9156. doi:10.1016/0040-4039(95)01961-G. ISSN   0040-4039.
  14. Venkataraman, Hemalatha; Cha, Jin K. (1989). "The scope and mechanism of rearrangement of 4,6-dialkoxy-2-pyrones". Tetrahedron Letters. 30 (27): 3509–3512. doi:10.1016/S0040-4039(00)99426-7. ISSN   0040-4039.
  15. Gridnev, Ilya D.; Tok, Oleg L.; Gridneva, Natalya A.; Bubnov, Yuri N.; Schreiner, Peter R. (1998). "Synthesis and Dynamic Properties of Cycloheptatrienyl(dipropyl)borane. Equilibrium with 7-Dipropylborylnorcaradiene". Journal of the American Chemical Society. 120 (5): 1034–1043. doi:10.1021/ja9724699. ISSN   0002-7863.
  16. Norris, R. K.; Sternhell, S. (1969). "2-Substituted and 2,6-disubstituted 1,4-benzoquinone 4-oximes ("p-nitrosophenols")". Aust. J. Chem. 22 (5): 935–970. doi:10.1071/CH9690935.
  17. Baldwin, J. E.; Norris, R. K. (1981). "Stereoelectronic control in organic chemistry: addition reactions of some 1,4-benzoquinone 4-(O-methyloximes)". J. Org. Chem. 46 (4): 697–703. doi:10.1021/jo00317a011.
  18. Perrin, C. L.; Engler, R. E. (1997). "Origin of Apparent Stereoelectronic Effects in Structure and Reactivity of Benzoquinone Monooximes". J. Org. Chem. 62 (3): 687–692. doi:10.1021/jo961386s. PMID   11671465.
  19. Winstein, S.; Shatavsky, M.; Norton, C.; Woodward, R. B. (1955). "7-Norbornenyl and 7-Norbornyl Cations". J. Am. Chem. Soc. 77 (15): 4183–4184. doi:10.1021/ja01620a078.
  20. 1 2 Mehta, Goverdhan; Uma, R. (2000). "Stereoelectronic Control in Diels−Alder Reaction of Dissymmetric 1,3-Dienes". Accounts of Chemical Research. 33 (5): 278–286. doi:10.1021/ar990123s. ISSN   0001-4842. PMID   10813872.
  21. Halterman, R. L.; McCarthy, B. A.; McEvoy, M. A. (1992). "Stereoelectronic control of facial selectivity in the Diels-Alder cycloaddition of sterically unbiased 5,5-diarylcyclopentadienes". J. Org. Chem. 57 (21): 5585–5589. doi:10.1021/jo00047a009.
  22. Shindo, M.; Sato, Y.; Shishido, J. (1998). "A highly stereoselective synthesis of tri- and tetrasubstituted olefins via ynolates". Tetrahedron Lett. 39 (27): 4857–4860. doi:10.1016/s0040-4039(98)00921-6.
  23. Murakami, M.; Hasegawa, M. (2004). "Synthesis and Thermal Ring Opening oftrans-3,4-Disilylcyclobutene". Angew. Chem. Int. Ed. 43 (37): 4874–4876. doi:10.1002/anie.200460144. PMID   15372640.
  24. Schlapbach, A.; Hoffmann, R. W. (2001). "(E)-α-Sulfonamidocrotylboronates as Reagents for the Stereoselective Homoaldol Synthesis". J. Org. Chem. 66 (2): 323–328. doi:10.1002/1099-0690(200101)2001:2<323::aid-ejoc323>3.0.co;2-a.
  25. Gorenstein, David G. (1987-10-01). "Stereoelectronic effects in biomolecules". Chemical Reviews. 87 (5): 1047–1077. doi:10.1021/cr00081a009. ISSN   0009-2665.
  26. Ardejani, Maziar S.; Noodleman, Louis; Powers, Evan T.; Kelly, Jeffery W. (2021-03-15). "Stereoelectronic effects in stabilizing protein– N -glycan interactions revealed by experiment and machine learning". Nature Chemistry: 1–8. doi:10.1038/s41557-021-00646-w. ISSN   1755-4349. PMC   8102341 .