Alpha effect

Last updated
Figure 1. Normal and a-nucleophiles Normal and alpha nucleophiles.png
Figure 1. Normal and α-nucleophiles

The alpha effect refers to the increased nucleophilicity of an atom due to the presence of an adjacent (alpha) atom with lone pair electrons. [2] This first atom does not necessarily exhibit increased basicity compared with a similar atom without an adjacent electron-donating atom, resulting in a deviation from the classical Brønsted-type reactivity-basicity relationship. In other words, the alpha effect refers to nucleophiles presenting higher nucleophilicity than the predicted value obtained from the Brønsted basicity. The representative examples would be high nucleophilicities of hydroperoxide (HO2) and hydrazine (N2H4). [3] The effect is now well established with numerous examples and became an important concept in mechanistic chemistry and biochemistry. [4] However, the origin of the effect is still controversial without a clear winner.

Contents

Background

Experiment

The effect was first observed by Jencks and Carriuolo in 1960 [5] [6] in a series of chemical kinetics experiments involving the reaction of the ester p-nitrophenyl acetate with a range of nucleophiles. Regular nucleophiles such as the fluoride anion, aniline, pyridine, ethylene diamine and the phenolate ion were found to have pseudo first order reaction rates corresponding to their basicity as measured by their pKa. Other nucleophiles however reacted much faster than expected based on this criterion alone. These include hydrazine, hydroxylamine, the hypochlorite ion and the hydroperoxide anion.

α-effect

In 1962, Edwards and Pearson (the latter of HSAB theory) introduced the phrase alpha effect for this anomaly. He offered the suggestion that the effect was caused by a transition state (TS) stabilization effect: on entering the TS the free electron pair on the nucleophile moves away from the nucleus, causing a partial positive charge which can be stabilized by an adjacent lone pair as for instance happens in any carbocation. [7]

Theory

Over the years, many additional theories have been put forward attempting to explain the effect.

Ground state destabilization

The ground state destabilization theory proposes that the electron-electron repulsion between the alpha lone-pair and nucleophilic electron pair destabilize each other by electronic repulsion (filled–filled orbital interaction) thereby decreasing the activation barrier by increasing the ground state energy and making it more reactive. This explains the higher reactivity of α-nucleophiles, however, this electronic mechanism should also increase the basicity and, therefore, cannot fully explain the α-effect. [4]

Transition state stabilization

Many explanations fall into this category. First, the secondary orbital interactions theory emphasized that electron-donating heteroatom in the α-position could contribute to increased orbital interaction with the substrate, which stabilizes the transition state (TS) and gives greater reactivity. [8] Second, the electron transfer (ET) mechanism presents that the heteroatom in the α position could stabilize the SN2 transition state which has a single electron transfer (free radical) character. [9] Other driving forces including the tighter transition state [10] and higher polarizability of α-nucleophiles, involvement of intramolecular catalysis also plays a role. Another in silico study did find a correlation between the alpha effect and the so-called deformation energy, which is the electronic energy required to bring the two reactants together in the transition state. [11]

Thermodynamic product stability

This explanation proposes that a stable product could contribute to the alpha effect, however, this factor could not be the sole factor. [4]

Solvent-induced effects

The alpha effect is also dependent on solvent but not in a predictable way: it can increase or decrease with solvent mix composition or even go through a maximum. [12] At least in some cases, the alpha effect has been observed to vanish if the reaction is conducted in the gas phase, leading some to conclude that it is primarily a solvation effect. [13] However, this explanation has limitations since similar alpha effects could be found in different solvent systems and also because the solvation affects both the basicity and the nucleophilicity of the nucleophile. [4]

Pauli repulsion and HOMO-LUMO overlap

Figure 2. SN2 reaction with the substrate, C2H5Cl Plot of reactivity and basicity.png
Figure 2. SN2 reaction with the substrate, C2H5Cl

In the recent article published in 2021, Hansen and Vermeeren proposed the two requirements for an α-nucleophile to present the α-effect. The two proposed requirements were (1) minimization of steric Pauli repulsion via small HOMO lobes on the nucleophilic center and (2) small HOMO-LUMO orbital energy gap that ensures orbital overlap with the substrate. It was proposed that both of these two requirements should be fulfilled to have an α-effect, otherwise, the nucleophiles would show no or inverse α-effect (Figure 2). In this recent work, the six normal nucleophiles (HO-, CH3O-, H2N-, CH3HN-, CH3S-, HS-) followed the Brønsted-type correlation. α-nucleophiles with O, HN, and S in the α position were classified into three groups according to their degree and pattern of deviation from the Brønsted-type correlation in SN2 reactions with the substrate, ethyl chloride (C2H5Cl) (Figure 3).

Figure 3. Plot of reactivity and basicity shows the (inverse)-a-effect Alpha and inverse alpha effect.png
Figure 3. Plot of reactivity and basicity shows the (inverse)-α-effect



First, the α-nucleophiles with downward deviation, in other words, higher reactivity shown considering the basicity or lower basicity given the reactivity, were grouped as nucleophiles showing α-effect. The second group had nucleophiles with small or no deviation from the line plotted by six normal nucleophiles. Lastly, the third group had nucleophiles showing inverse α-effect, meaning that they are above the plotted line or have less reactivity considering their high basicity. Relative density functional theory, activation strain model, energy decomposition analysis, and Kohn-Sham molecular orbital analysis the three groups had a distinction in HOMO lobes and HOMO-LUMO gaps.

To elaborate on the first requirement, the electronegative heteroatom reduces the electron density of the atom that attacks the nucleophile making the HOMO lobe smaller. This minimizes the Pauli repulsion between the substrate and the nucleophile. Nonetheless, these small HOMO lobes don't mean less orbital interaction than the parent normal nucleophile. This is because α-nucleophiles showing the α-effect have smaller HOMO(nucleophile)-LUMO(substrate) gap, in other words, high HOMO energy level that allows more orbital interaction.

Examples of α-nucleophiles with α-effects are shown in Figure 4. The α-nucleophiles have smaller HOMO lobes than the parent normal nucleophile.

Figure 4. Normal and a-nucleophiles (smaller HOMO lobes) Alpha nucleophiles.png
Figure 4. Normal and α-nucleophiles (smaller HOMO lobes)

Examples of α-nucleophiles with α-effect and inverse α-effect are shown in Figure 5.

Figure 5. a-effect and inverse a-effect A-effect and inverse a-effect.png
Figure 5. α-effect and inverse α-effect

See also

Related Research Articles

In chemistry, a nucleophile is a chemical species that forms bonds by donating an electron pair. All molecules and ions with a free pair of electrons or at least one pi bond can act as nucleophiles. Because nucleophiles donate electrons, they are Lewis bases.

In chemistry, a nucleophilic substitution is a class of chemical reactions in which an electron-rich chemical species replaces a functional group within another electron-deficient molecule. The molecule that contains the electrophile and the leaving functional group is called the substrate.

<span class="mw-page-title-main">Elimination reaction</span> Reaction where 2 substituents are removed from a molecule in a 1 or 2 step mechanism

An elimination reaction is a type of organic reaction in which two substituents are removed from a molecule in either a one- or two-step mechanism. The one-step mechanism is known as the E2 reaction, and the two-step mechanism is known as the E1 reaction. The numbers refer not to the number of steps in the mechanism, but rather to the kinetics of the reaction: E2 is bimolecular (second-order) while E1 is unimolecular (first-order). In cases where the molecule is able to stabilize an anion but possesses a poor leaving group, a third type of reaction, E1CB, exists. Finally, the pyrolysis of xanthate and acetate esters proceed through an "internal" elimination mechanism, the Ei mechanism.

The unimolecular nucleophilic substitution (SN1) reaction is a substitution reaction in organic chemistry. The Hughes-Ingold symbol of the mechanism expresses two properties—"SN" stands for "nucleophilic substitution", and the "1" says that the rate-determining step is unimolecular. Thus, the rate equation is often shown as having first-order dependence on the substrate and zero-order dependence on the nucleophile. This relationship holds for situations where the amount of nucleophile is much greater than that of the intermediate. Instead, the rate equation may be more accurately described using steady-state kinetics. The reaction involves a carbocation intermediate and is commonly seen in reactions of secondary or tertiary alkyl halides under strongly basic conditions or, under strongly acidic conditions, with secondary or tertiary alcohols. With primary and secondary alkyl halides, the alternative SN2 reaction occurs. In inorganic chemistry, the SN1 reaction is often known as the dissociative substitution. This dissociation pathway is well-described by the cis effect. A reaction mechanism was first proposed by Christopher Ingold et al. in 1940. This reaction does not depend much on the strength of the nucleophile, unlike the SN2 mechanism. This type of mechanism involves two steps. The first step is the ionization of alkyl halide in the presence of aqueous acetone or ethyl alcohol. This step provides a carbocation as an intermediate.

<span class="mw-page-title-main">Leaving group</span> Atom(s) which detach from the substrate during a chemical reaction

In chemistry, a leaving group is defined by the IUPAC as an atom or group of atoms that detaches from the main or residual part of a substrate during a reaction or elementary step of a reaction. However, in common usage, the term is often limited to a fragment that departs with a pair of electrons in heterolytic bond cleavage. In this usage, a leaving group is a less formal but more commonly used synonym of the term nucleofuge. In this context, leaving groups are generally anions or neutral species, departing from neutral or cationic substrates, respectively, though in rare cases, cations leaving from a dicationic substrate are also known.

S<sub>N</sub>2 reaction Substitution reaction where bonds are broken and formed simultaneously

Bimolecular nucleophilic substitution (SN2) is a type of reaction mechanism that is common in organic chemistry. In the SN2 reaction, a strong nucleophile forms a new bond to an sp3-hybridised carbon atom via a backside attack, all while the leaving group detaches from the reaction center in a concerted fashion.

<span class="mw-page-title-main">Organolithium reagent</span> Chemical compounds containing C–Li bonds

In organometallic chemistry, organolithium reagents are chemical compounds that contain carbon–lithium (C–Li) bonds. These reagents are important in organic synthesis, and are frequently used to transfer the organic group or the lithium atom to the substrates in synthetic steps, through nucleophilic addition or simple deprotonation. Organolithium reagents are used in industry as an initiator for anionic polymerization, which leads to the production of various elastomers. They have also been applied in asymmetric synthesis in the pharmaceutical industry. Due to the large difference in electronegativity between the carbon atom and the lithium atom, the C−Li bond is highly ionic. Owing to the polar nature of the C−Li bond, organolithium reagents are good nucleophiles and strong bases. For laboratory organic synthesis, many organolithium reagents are commercially available in solution form. These reagents are highly reactive, and are sometimes pyrophoric.

A substitution reaction is a chemical reaction during which one functional group in a chemical compound is replaced by another functional group. Substitution reactions are of prime importance in organic chemistry. Substitution reactions in organic chemistry are classified either as electrophilic or nucleophilic depending upon the reagent involved, whether a reactive intermediate involved in the reaction is a carbocation, a carbanion or a free radical, and whether the substrate is aliphatic or aromatic. Detailed understanding of a reaction type helps to predict the product outcome in a reaction. It also is helpful for optimizing a reaction with regard to variables such as temperature and choice of solvent.

The 1,3-dipolar cycloaddition is a chemical reaction between a 1,3-dipole and a dipolarophile to form a five-membered ring. The earliest 1,3-dipolar cycloadditions were described in the late 19th century to the early 20th century, following the discovery of 1,3-dipoles. Mechanistic investigation and synthetic application were established in the 1960s, primarily through the work of Rolf Huisgen. Hence, the reaction is sometimes referred to as the Huisgen cycloaddition. 1,3-dipolar cycloaddition is an important route to the regio- and stereoselective synthesis of five-membered heterocycles and their ring-opened acyclic derivatives. The dipolarophile is typically an alkene or alkyne, but can be other pi systems. When the dipolarophile is an alkyne, aromatic rings are generally produced.

<span class="mw-page-title-main">Nucleophilic aromatic substitution</span> Chemical reaction mechanism

A nucleophilic aromatic substitution is a substitution reaction in organic chemistry in which the nucleophile displaces a good leaving group, such as a halide, on an aromatic ring. Aromatic rings are usually nucleophilic, but some aromatic compounds do undergo nucleophilic substitution. Just as normally nucleophilic alkenes can be made to undergo conjugate substitution if they carry electron-withdrawing substituents, so normally nucleophilic aromatic rings also become electrophilic if they have the right substituents.

In organic chemistry, neighbouring group participation has been defined by the International Union of Pure and Applied Chemistry (IUPAC) as the interaction of a reaction centre with a lone pair of electrons in an atom or the electrons present in a sigma or pi bond contained within the parent molecule but not conjugated with the reaction centre. When NGP is in operation it is normal for the reaction rate to be increased. It is also possible for the stereochemistry of the reaction to be abnormal when compared with a normal reaction. While it is possible for neighbouring groups to influence many reactions in organic chemistry this page is limited to neighbouring group effects seen with carbocations and SN2 reactions.

<span class="mw-page-title-main">Shi epoxidation</span>

The Shi epoxidation is a chemical reaction described as the asymmetric epoxidation of alkenes with oxone and a fructose-derived catalyst (1). This reaction is thought to proceed via a dioxirane intermediate, generated from the catalyst ketone by oxone. The addition of the sulfate group by the oxone facilitates the formation of the dioxirane by acting as a good leaving group during ring closure. It is notable for its use of a non-metal catalyst and represents an early example of organocatalysis.

<span class="mw-page-title-main">Asymmetric induction</span> Preferential formation of one chiral isomer over another in a chemical reaction

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.

<span class="mw-page-title-main">Vinylogy</span> Transmission of electronic effects through a system of conjugated chemical bonds

In organic chemistry, vinylogy is the transmission of electronic effects through a conjugated organic bonding system. The concept was introduced in 1926 by Ludwig Claisen to explain the acidic properties of formylacetone and related ketoaldehydes. Formylacetone, technically CH3(C=O)CH2CH=O, only exists in the ionized form CH3(C−O)=CH−CH=O or CH3(C=O)−CH=CH−O. Its adjectival form, vinylogous, is used to describe functional groups in which the standard moieties of the group are separated by a carbon–carbon double bond.

<span class="mw-page-title-main">Flippin–Lodge angle</span>

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.

Arrow pushing or electron pushing is a technique used to describe the progression of organic chemistry reaction mechanisms. It was first developed by Sir Robert Robinson. In using arrow pushing, "curved arrows" or "curly arrows" are drawn on the structural formulae of reactants in a chemical equation to show the reaction mechanism. The arrows illustrate the movement of electrons as bonds between atoms are broken and formed. It is important to note that arrow pushing never directly show the movement of atoms; it is used to show the movement of electron density, which indirectly shows the movement of atoms themselves. Arrow pushing is also used to describe how positive and negative charges are distributed around organic molecules through resonance. It is important to remember, however, that arrow pushing is a formalism and electrons do not move around so neatly and discretely in reality.

In chemistry, solvent effects are the influence of a solvent on chemical reactivity or molecular associations. Solvents can have an effect on solubility, stability and reaction rates and choosing the appropriate solvent allows for thermodynamic and kinetic control over a chemical reaction.

The Edwards equation in organic chemistry is a two-parameter equation for correlating nucleophilic reactivity, as defined by relative rate constants, with the basicity of the nucleophile and its polarizability. This equation was first developed by John O. Edwards in 1954 and later revised based on additional work in 1956.

A phosphetane is a 4-membered organophosphorus heterocycle. The parent phosphetane molecule, which has the formula C3H7P, is one atom larger than phosphiranes, one smaller than phospholes, and is the heavy-atom analogue of azetidines. The first known phosphetane synthesis was reported in 1957 by Kosolapoff and Struck, but the method was both inefficient and hard to reproduce, with yields rarely exceeding 1%. A far more efficient method was reported in 1962 by McBride, whose method allowed for the first studies into the physical and chemical properties of phosphetanes. Phosphetanes are a well understood class of molecules that have found broad applications as chemical building blocks, reagents for organic/inorganic synthesis, and ligands in coordination chemistry.

Negative hyperconjugation is a theorized phenomenon in organosilicon compounds, in which hyperconjugation stabilizes or destabilizes certain accumulations of positive charge. The phenomenon explains corresponding peculiarities in the stereochemistry and rate of hydrolysis.

References

  1. 1 2 3 4 5 Hansen, Thomas; Vermeeren, Pascal; Bickelhaupt, F. Matthias; Hamlin, Trevor A. (2021-07-26). "Origin of the α-Effect in SN2 Reactions". Angewandte Chemie International Edition. 60 (38): 20840–20848. doi: 10.1002/anie.202106053 . ISSN   1433-7851. PMC   8518820 . PMID   34087047.
  2. Chemical Reactivity . 14 July 2006. Michigan State University. 27 Jul 2006 <>.
  3. Smith, Michael B.; March, Jerry (2007), Advanced Organic Chemistry: Reactions, Mechanisms, and Structure (6th ed.), New York: Wiley-Interscience, p. 495, ISBN   978-0-471-72091-1
  4. 1 2 3 4 Fina, Nick J.; Edwards, John O. (January 1973). "The alpha effect. A review". International Journal of Chemical Kinetics. 5 (1): 1–26. doi:10.1002/kin.550050102. ISSN   0538-8066.
  5. William P. Jencks; Joan Carriuolo (1960). "Reactivity of Nucleophilic Reagents toward Esters". Journal of the American Chemical Society. 82 (7): 1778–86. doi:10.1021/ja01492a058.
  6. William P. Jencks; Joan Carriuolo (1960). "General Base Catalysis of the Aminolysis of Phenyl Acetate". Journal of the American Chemical Society. 82 (3): 675–81. doi:10.1021/ja01488a044.
  7. John O. Edwards; Ralph G. Pearson (1962). "The Factors Determining Nucleophilic Reactivities". Journal of the American Chemical Society. 84: 16–24. doi:10.1021/ja00860a005.
  8. Liebman, Joel F.; Pollack, Ralph M. (September 1973). "Aromatic transition states and the .alpha. effect". The Journal of Organic Chemistry. 38 (19): 3444–3445. doi:10.1021/jo00959a059. ISSN   0022-3263.
  9. Hoz, Shmaryahu (October 1982). "The .alpha. effect: on the origin of transition-state stabilization". The Journal of Organic Chemistry. 47 (18): 3545–3547. doi:10.1021/jo00139a033. ISSN   0022-3263.
  10. Fountain, K. R.; Felkerson, Cassie J.; Driskell, Jeremy D.; Lamp, Brian D. (2003-03-01). "The α-Effect in Methyl Transfers from S -Methyldibenzothiophenium Fluoroborate to Substituted N -Methylbenzohydroxamates". The Journal of Organic Chemistry. 68 (5): 1810–1814. doi:10.1021/jo0206263. ISSN   0022-3263. PMID   12608795.
  11. Ren, Y; Yamataka, H (Jul 2007). "The alpha-effect in gas-phase SN2 reactions: existence and the origin of the effect". The Journal of Organic Chemistry. 72 (15): 5660–7. doi:10.1021/jo070650m. ISSN   0022-3263. PMID   17590049.
  12. Buncel, Erwin; Um, Ik-Hwan (2004). "The α-effect and its modulation by solvent". Tetrahedron. 60 (36): 7801. doi:10.1016/j.tet.2004.05.006.
  13. A., Carroll, Felix (2010). Perspectives on structure and mechanism in organic chemistry (2nd ed.). Hoboken, N.J.: John Wiley. ISBN   9780470276105. OCLC   286483846.{{cite book}}: CS1 maint: multiple names: authors list (link)