Electrophile

Last updated

In chemistry, an electrophile is a chemical species that forms bonds with nucleophiles by accepting an electron pair. [1] Because electrophiles accept electrons, they are Lewis acids. [2] Most electrophiles are positively charged, have an atom that carries a partial positive charge, or have an atom that does not have an octet of electrons.

Contents

Electrophiles mainly interact with nucleophiles through addition and substitution reactions. Frequently seen electrophiles in organic syntheses include cations such as H+ and NO+, polarized neutral molecules such as HCl, alkyl halides, acyl halides, and carbonyl compounds, polarizable neutral molecules such as Cl2 and Br2, oxidizing agents such as organic peracids, chemical species that do not satisfy the octet rule such as carbenes and radicals, and some Lewis acids such as BH3 and DIBAL.

Organic chemistry

Addition of halogens

These occur between alkenes and electrophiles, often halogens as in halogen addition reactions. Common reactions include use of bromine water to titrate against a sample to deduce the number of double bonds present. For example, ethene + bromine1,2-dibromoethane:

C2H4 + Br2 → BrCH2CH2Br

This takes the form of 3 main steps shown below; [3]

Electrophilic addition of Br2.png
  1. Forming of a π-complex
    The electrophilic Br-Br molecule interacts with electron-rich alkene molecule to form a π-complex 1.
  2. Forming of a three-membered bromonium ion
    The alkene is working as an electron donor and bromine as an electrophile. The three-membered bromonium ion 2 consisted of two carbon atoms and a bromine atom forms with a release of Br.
  3. Attacking of bromide ion
    The bromonium ion is opened by the attack of Br from the back side. This yields the vicinal dibromide with an antiperiplanar configuration. When other nucleophiles such as water or alcohol are existing, these may attack 2 to give an alcohol or an ether.

This process is called AdE2 mechanism ("addition, electrophilic, second-order"). Iodine (I2), chlorine (Cl2), sulfenyl ion (RS+), mercury cation (Hg2+), and dichlorocarbene (:CCl2) also react through similar pathways. The direct conversion of 1 to 3 will appear when the Br is large excess in the reaction medium. A β-bromo carbenium ion intermediate may be predominant instead of 3 if the alkene has a cation-stabilizing substituent like phenyl group. There is an example of the isolation of the bromonium ion 2. [4]

Addition of hydrogen halides

Hydrogen halides such as hydrogen chloride (HCl) adds to alkenes to give alkyl halides in hydrohalogenation. For example, the reaction of HCl with ethylene furnishes chloroethane. The reaction proceeds with a cation intermediate, being different from the above halogen addition. An example is shown below:

Electrophilic addition of HCl.png
  1. Proton (H+) adds (by working as an electrophile) to one of the carbon atoms on the alkene to form cation 1.
  2. Chloride ion (Cl) combines with the cation 1 to form the adducts 2 and 3.

In this manner, the stereoselectivity of the product, that is, from which side Cl will attack relies on the types of alkenes applied and conditions of the reaction. At least, which of the two carbon atoms will be attacked by H+ is usually decided by Markovnikov's rule. Thus, H+ attacks the carbon atom that carries fewer substituents so as the more stabilized carbocation (with the more stabilizing substituents) will form.

This is another example of an AdE2 mechanism. [5] Hydrogen fluoride (HF) and hydrogen iodide (HI) react with alkenes in a similar manner, and Markovnikov-type products will be given. Hydrogen bromide (HBr) also takes this pathway, but sometimes a radical process competes and a mixture of isomers may form. Although introductory textbooks seldom mentions this alternative, [6] the AdE2 mechanism is generally competitive with the AdE3 mechanism (described in more detail for alkynes, below), in which transfer of the proton and nucleophilic addition occur in a concerted manner. The extent to which each pathway contributes depends on the several factors like the nature of the solvent (e.g., polarity), nucleophilicity of the halide ion, stability of the carbocation, and steric effects. As brief examples, the formation of a sterically unencumbered, stabilized carbocation favors the AdE2 pathway, while a more nucleophilic bromide ion favors the AdE3 pathway to a greater extent compared to reactions involving the chloride ion. [7]

In the case of dialkyl-substituted alkynes (e.g., 3-hexyne), the intermediate vinyl cation that would result from this process is highly unstable. In such cases, the simultaneous protonation (by HCl) and attack of the alkyne by the nucleophile (Cl) is believed to take place. This mechanistic pathway is known by the Ingold label AdE3 ("addition, electrophilic, third-order"). Because the simultaneous collision of three chemical species in a reactive orientation is improbable, the termolecular transition state is believed to be reached when the nucleophile attacks a reversibly-formed weak association of the alkyne and HCl. Such a mechanism is consistent with the predominantly anti addition (>15:1 anti:syn for the example shown) of the hydrochlorination product and the termolecular rate law, Rate = k[alkyne][HCl]2. [8] [9] In support of the proposed alkyne-HCl association, a T-shaped complex of an alkyne and HCl has been characterized crystallographically. [10]

Ade3 reaction.png

In contrast, phenylpropyne reacts by the AdE2ip ("addition, electrophilic, second-order, ion pair") mechanism to give predominantly the syn product (~10:1 syn:anti). In this case, the intermediate vinyl cation is formed by addition of HCl because it is resonance-stabilized by the phenyl group. Nevertheless, the lifetime of this high energy species is short, and the resulting vinyl cation-chloride anion ion pair immediately collapses, before the chloride ion has a chance to leave the solvent shell, to give the vinyl chloride. The proximity of the anion to the side of the vinyl cation where the proton was added is used to rationalize the observed predominance of syn addition. [7]

Ade2ip.png


Hydration

One of the more complex hydration reactions utilises sulfuric acid as a catalyst. This reaction occurs in a similar way to the addition reaction but has an extra step in which the OSO3H group is replaced by an OH group, forming an alcohol:

C2H4 + H2O → C2H5OH

As can be seen, the H2SO4 does take part in the overall reaction, however it remains unchanged so is classified as a catalyst.

This is the reaction in more detail:

Electrophilic reaction of sulfuric acid with ethene.png
  1. The H–OSO3H molecule has a δ+ charge on the initial H atom. This is attracted to and reacts with the double bond in the same way as before.
  2. The remaining (negatively charged) OSO3H ion then attaches to the carbocation, forming ethyl hydrogensulphate (upper way on the above scheme).
  3. When water (H2O) is added and the mixture heated, ethanol (C2H5OH) is produced. The "spare" hydrogen atom from the water goes into "replacing" the "lost" hydrogen and, thus, reproduces sulfuric acid. Another pathway in which water molecule combines directly to the intermediate carbocation (lower way) is also possible. This pathway become predominant when aqueous sulfuric acid is used.

Overall, this process adds a molecule of water to a molecule of ethene.

This is an important reaction in industry, as it produces ethanol, whose purposes include fuels and starting material for other chemicals.

Chiral derivatives

Many electrophiles are chiral and optically stable. Typically chiral electrophiles are also optically pure.

One such reagent is the fructose-derived organocatalyst used in the Shi epoxidation. [11] The catalyst can accomplish highly enantioselective epoxidations of trans-disubstituted and trisubstituted alkenes. The Shi catalyst, a ketone, is oxidized by stoichiometric oxone to the active dioxirane form before proceeding in the catalytic cycle.

Use of a chiral oxaziridine for asymmetric synthesis. 18 fig. 1.png
Use of a chiral oxaziridine for asymmetric synthesis.

Oxaziridines such as chiral N-sulfonyloxaziridines effect enantioselective ketone alpha oxidation en route to the AB-ring segments of various natural products, including γ-rhodomycionone and α-citromycinone. [12]

Polymer-bound chiral selenium electrophiles effect asymmetric selenenylation reactions. [13] The reagents are aryl selenenyl bromides, and they were first developed for solution phase chemistry and then modified for solid phase bead attachment via an aryloxy moiety. The solid-phase reagents were applied toward the selenenylation of various alkenes with good enantioselectivities. The products can be cleaved from the solid support using organotin hydride reducing agents. Solid-supported reagents offers advantages over solution phase chemistry due to the ease of workup and purification.

Electrophilicity scale

Electrophilicity index
Fluorine 3.86
Chlorine 3.67
Bromine 3.40
Iodine 3.09
Hypochlorite 2.52
Sulfur dioxide 2.01
Carbon disulfide 1.64
Benzene 1.45
Sodium 0.88
Some selected values [14] (no dimensions)

Several methods exist to rank electrophiles in order of reactivity [15] and one of them is devised by Robert Parr [14] with the electrophilicity index ω given as:

with the electronegativity and chemical hardness. This equation is related to the classical equation for electrical power:

where is the resistance (Ohm or Ω) and is voltage. In this sense the electrophilicity index is a kind of electrophilic power. Correlations have been found between electrophilicity of various chemical compounds and reaction rates in biochemical systems and such phenomena as allergic contact dermititis.

An electrophilicity index also exists for free radicals. [16] Strongly electrophilic radicals such as the halogens react with electron-rich reaction sites, and strongly nucleophilic radicals such as the 2-hydroxypropyl-2-yl and tert-butyl radical react with a preference for electron-poor reaction sites.

Superelectrophiles

Superelectrophiles are defined as cationic electrophilic reagents with greatly enhanced reactivities in the presence of superacids. These compounds were first described by George A. Olah. [17] Superelectrophiles form as a doubly electron deficient superelectrophile by protosolvation of a cationic electrophile. As observed by Olah, a mixture of acetic acid and boron trifluoride is able to remove a hydride ion from isobutane when combined with hydrofluoric acid via the formation of a superacid from BF3 and HF. The responsible reactive intermediate is the [CH3CO2H3]2+ dication. Likewise, methane can be nitrated to nitromethane with nitronium tetrafluoroborate NO+
2
BF
4
only in presence of a strong acid like fluorosulfuric acid via the protonated nitronium dication.

In gitionic (gitonic) superelectrophiles, charged centers are separated by no more than one atom, for example, the protonitronium ion O=N+=O+—H (a protonated nitronium ion). And, in distonic superelectrophiles, they are separated by 2 or more atoms, for example, in the fluorination reagent F-TEDA-BF4. [18]

See also

Related Research Articles

<span class="mw-page-title-main">Chemical reaction</span> Process that results in the interconversion of chemical species

A chemical reaction is a process that leads to the chemical transformation of one set of chemical substances to another. Classically, chemical reactions encompass changes that only involve the positions of electrons in the forming and breaking of chemical bonds between atoms, with no change to the nuclei, and can often be described by a chemical equation. Nuclear chemistry is a sub-discipline of chemistry that involves the chemical reactions of unstable and radioactive elements where both electronic and nuclear changes can occur.

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.

<span class="mw-page-title-main">Haloalkane</span> Group of chemical compounds derived from alkanes containing one or more halogens

The haloalkanes are alkanes containing one or more halogen substituents. They are a subset of the general class of halocarbons, although the distinction is not often made. Haloalkanes are widely used commercially. They are used as flame retardants, fire extinguishants, refrigerants, propellants, solvents, and pharmaceuticals. Subsequent to the widespread use in commerce, many halocarbons have also been shown to be serious pollutants and toxins. For example, the chlorofluorocarbons have been shown to lead to ozone depletion. Methyl bromide is a controversial fumigant. Only haloalkanes that contain chlorine, bromine, and iodine are a threat to the ozone layer, but fluorinated volatile haloalkanes in theory may have activity as greenhouse gases. Methyl iodide, a naturally occurring substance, however, does not have ozone-depleting properties and the United States Environmental Protection Agency has designated the compound a non-ozone layer depleter. For more information, see Halomethane. Haloalkane or alkyl halides are the compounds which have the general formula "RX" where R is an alkyl or substituted alkyl group and X is a halogen.

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.

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.

In organic chemistry, Markovnikov's rule or Markownikoff's rule describes the outcome of some addition reactions. The rule was formulated by Russian chemist Vladimir Markovnikov in 1870.

In organic chemistry, the oxymercuration reaction is an electrophilic addition reaction that transforms an alkene into a neutral alcohol. In oxymercuration, the alkene reacts with mercuric acetate in aqueous solution to yield the addition of an acetoxymercury group and a hydroxy group across the double bond. Carbocations are not formed in this process and thus rearrangements are not observed. The reaction follows Markovnikov's rule and it is an anti addition.

<span class="mw-page-title-main">Carbocation</span> Ion with a positively charged carbon atom

A carbocation is an ion with a positively charged carbon atom. Among the simplest examples are the methenium CH+
3
, methanium CH+
5
and vinyl C
2
H+
3
cations. Occasionally, carbocations that bear more than one positively charged carbon atom are also encountered.

<span class="mw-page-title-main">Beckmann rearrangement</span> Chemical rearrangement

The Beckmann rearrangement, named after the German chemist Ernst Otto Beckmann (1853–1923), is a rearrangement of an oxime functional group to substituted amides. The rearrangement has also been successfully performed on haloimines and nitrones. Cyclic oximes and haloimines yield lactams.

In organic chemistry, a nucleophilic addition reaction is an addition reaction where a chemical compound with an electrophilic double or triple bond reacts with a nucleophile, such that the double or triple bond is broken. Nucleophilic additions differ from electrophilic additions in that the former reactions involve the group to which atoms are added accepting electron pairs, whereas the latter reactions involve the group donating electron pairs.

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.

<span class="mw-page-title-main">Hydrohalogenation</span> Electrophilic addition of hydrogen halides to alkenes

A hydrohalogenation reaction is the electrophilic addition of hydrogen halides like hydrogen chloride or hydrogen bromide to alkenes to yield the corresponding haloalkanes.

<span class="mw-page-title-main">Bamford–Stevens reaction</span> Synthesis of alkenes by base-catalysed decomposition of tosylhydrazones

The Bamford–Stevens reaction is a chemical reaction whereby treatment of tosylhydrazones with strong base gives alkenes. It is named for the British chemist William Randall Bamford and the Scottish chemist Thomas Stevens Stevens (1900–2000). The usage of aprotic solvents gives predominantly Z-alkenes, while protic solvent gives a mixture of E- and Z-alkenes. As an alkene-generating transformation, the Bamford–Stevens reaction has broad utility in synthetic methodology and complex molecule synthesis.

<span class="mw-page-title-main">Carbenium ion</span> Class of ions

A carbenium ion is a positive ion with the structure RR′R″C+, that is, a chemical species with carbon atom having three covalent bonds, and it bears a +1 formal charge. But IUPAC confuses coordination number with valence, incorrectly considering carbon in carbenium as trivalent.

<span class="mw-page-title-main">Halonium ion</span> Any onium ion containing a halogen atom carrying a positive charge

A halonium ion is any onium ion containing a halogen atom carrying a positive charge. This cation has the general structure R−+X−R′ where X is any halogen and no restrictions on R, this structure can be cyclic or an open chain molecular structure. Halonium ions formed from fluorine, chlorine, bromine, and iodine are called fluoronium, chloronium, bromonium, and iodonium, respectively. The 3-membered cyclic variety commonly proposed as intermediates in electrophilic halogenation may be called haliranium ions, using the Hantzsch-Widman nomenclature system.

In chemistry, a reaction intermediate, or intermediate, is a molecular entity arising within the sequence of a stepwise chemical reaction. It is formed as the reaction product of an elementary step, from the reactants and/or preceding intermediates, but is consumed in a later step. It does not appear in the chemical equation for the overall reaction.

An insertion reaction is a chemical reaction where one chemical entity interposes itself into an existing bond of typically a second chemical entity e.g.:

Electrophilic aromatic substitution is an organic reaction in which an atom that is attached to an aromatic system is replaced by an electrophile. Some of the most important electrophilic aromatic substitutions are aromatic nitration, aromatic halogenation, aromatic sulfonation, alkylation and acylation Friedel–Crafts reaction.

<span class="mw-page-title-main">Carbonyl α-substitution reactions</span>

Alpha-substitution reactions occur at the position next to the carbonyl group, the α-position, and involve the substitution of an α hydrogen atom by an electrophile, E, through either an enol or enolate ion intermediate.

Hydrogen-bridged cations are a type of charged species in which a hydrogen atom is simultaneously bonded to two atoms through partial sigma bonds. While best observable in the presence of superacids at room temperature, spectroscopic evidence has suggested that hydrogen-bridged cations exist in ordinary solvents. These ions have been the subject of debate as they constitute a type of charged species of uncertain electronic structure.

References

  1. "Nucleophiles and Electrophiles". butane.chem.uiuc.edu. Archived from the original on 2020-08-01. Retrieved 2020-09-21.
  2. "Electrophile | chemistry". Encyclopedia Britannica. Retrieved 2020-09-21.
  3. Lenoir, D.; Chiappe, C. (2003). "What is the Nature of the First-Formed Intermediates in the Electrophilic Halogenation of Alkenes, Alkynes, and Allenes?". Chem. Eur. J. 9 (5): 1036–1044. doi:10.1002/chem.200390097. PMID   12596140.
  4. Brown, R. S. (1997). "Investigation of the Early Steps in Electrophilic Bromination through the Study of the Reaction with Sterically Encumbered Olefins". Acc. Chem. Res. 30 (3): 131–137. doi:10.1021/ar960088e.
  5. In analogy to aromatic substitution, this process has also been termed an A-SE2 mechanism.
  6. Vollhardt, K. Peter C.; Schore, Neil Eric (January 2014). Organic chemistry : structure and function (7th ed.). New York, NY. ISBN   978-1-4641-2027-5. OCLC   866584251.{{cite book}}: CS1 maint: location missing publisher (link)
  7. 1 2 H., Lowry, Thomas (1987). Mechanism and theory in organic chemistry . Richardson, Kathleen Schueller. (3rd ed.). New York: Harper & Row. ISBN   978-0060440848. OCLC   14214254.{{cite book}}: CS1 maint: multiple names: authors list (link)
  8. Fahey, Robert C.; Lee, Do-Jae. (April 1968). "Polar additions to olefins and acetylenes. V. Bimolecular and termolecular mechanisms in the hydrochlorination of acetylenes". Journal of the American Chemical Society. 90 (8): 2124–2131. doi:10.1021/ja01010a034. ISSN   0002-7863.
  9. 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)
  10. Mootz, Dietrich; Deeg, Axel (July 1992). "2-Butyne and hydrogen chloride cocrystallized: solid-state geometry of Cl-H.cntdot..cntdot..cntdot..pi. hydrogen bonding to the carbon-carbon triple bond". Journal of the American Chemical Society. 114 (14): 5887–5888. doi:10.1021/ja00040a077. ISSN   0002-7863.
  11. Wang, Z.; Tu, Y.; Frohn, M.; Zhang, J.; Shi, Y. (1997). "An Efficient Catalytic Asymmetric Epoxidation Method". J. Am. Chem. Soc. 119 (46): 11224–11235. doi:10.1021/ja972272g.
  12. Davis, F. A.; Kumar, A.; Chen, B. C. (1991). "Chemistry of oxaziridines. 16. A short, highly enantioselective synthesis of the AB-ring segments of γ-rhodomycionone and α-citromycinone using (+)-[(8,8-dimethoxycamphoryl)sulfonyl]oxaziridine". J. Org. Chem. 56 (3): 1143–1145. doi:10.1021/jo00003a042.
  13. Uehlin, L.; Wirth, T. (2001). "Novel Polymer-Bound Chiral Selenium Electrophiles". Org. Lett. 3 (18): 2931–2933. doi:10.1021/ol0164435. PMID   11529793.
  14. 1 2 Parr, R. G.; Szentpaly, L. v.; Liu, S. (1999). "Electrophilicity Index". J. Am. Chem. Soc. 121 (9): 1922–1924. doi:10.1021/ja983494x.
  15. Chattaraj, P. K.; Sarkar, U.; Roy, D. R. (2006). "Electrophilicity Index". Chem. Rev. 106 (6): 2065–2091. doi:10.1021/cr040109f. PMID   16771443.
  16. De Vleeschouwer, Freija; Speybroeck, Veronique Van; Waroquier, Michel; Geerlings, Paul; De Proft, Frank (2007). "Electrophilicity and Nucleophilicity Index for Radicals". Org. Lett. 9 (14): 2721–2724. doi:10.1021/ol071038k. PMID   17559221.
  17. Olah, George A.; Germain, Alain; Lin, Henry C.; Forsyth, David A. (1975). "Electrophilic reactions at single bonds. XVIII. Indication of protosolvated de facto substituting agents in the reactions of alkanes with acetylium and nitronium ions in superacidic media". J. Am. Chem. Soc. 97 (10): 2928–2929. doi:10.1021/ja00843a067.
  18. Solingapuram Sai, Kiran Kumar; Gilbert, Thomas M.; Klumpp, Douglas A. (2007). "Knorr Cyclizations and Distonic Superelectrophiles". J. Org. Chem. 72 (25): 9761–9764. doi:10.1021/jo7013092. PMID   17999519.
  19. Lin King, John V.; Emrick, Joshua J.; Kelly, Mark J. S.; Herzig, Volker; King, Glenn F.; Medzihradszky, Katalin F.; Julius, David (5 September 2019). "A Cell-Penetrating Scorpion Toxin Enables Mode-Specific Modulation of TRPA1 and Pain". Cell. 178 (6): 1362–1374.e16. doi:10.1016/j.cell.2019.07.014. ISSN   1097-4172. PMC   6731142 . PMID   31447178.
  20. Zhao, Jianhua; Lin King, John V.; Paulsen, Candice E.; Cheng, Yifan; Julius, David (2020-07-08). "Irritant-evoked activation and calcium modulation of the TRPA1 receptor". Nature. 585 (7823): 141–145. doi:10.1038/s41586-020-2480-9. ISSN   1476-4687. PMC   7483980 . PMID   32641835.