Nucleophilic aromatic substitution

Last updated

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.

Contents

Aromatic nucleophilic substitution Aromatic nucleophilic substitution.svg
Aromatic nucleophilic substitution

This reaction differs from a common SN2 reaction, because it happens at a trigonal carbon atom (sp2 hybridization). The mechanism of SN2 reaction does not occur due to steric hindrance of the benzene ring. In order to attack the C atom, the nucleophile must approach in line with the C-LG (leaving group) bond from the back, where the benzene ring lies. It follows the general rule for which SN2 reactions occur only at a tetrahedral carbon atom.

The SN1 mechanism is possible but very unfavourable unless the leaving group is an exceptionally good one. It would involve the unaided loss of the leaving group and the formation of an aryl cation. In the SN1 reactions all the cations employed as intermediates were planar with an empty p orbital. This cation is planar but the p orbital is full (it is part of the aromatic ring) and the empty orbital is an sp2 orbital outside the ring. [1]

Nucleophilic aromatic substitution mechanisms

Aromatic rings undergo nucleophilic substitution by several pathways.

  1. SNAr (addition-elimination) mechanism
    SNAr mechanism.svg
  2. aromatic SN1 mechanism encountered with diazonium salts
    Aromatic SN1 mechanism.svg
  3. benzyne mechanism (E1cB-AdN)
    Substitution via benzyne.svg
  4. free radical SRN1 mechanism
  5. ANRORC mechanism
  6. Vicarious nucleophilic substitution

The SNAr mechanism is the most important of these. Electron withdrawing groups activate the ring towards nucleophilic attack. For example if there are nitro functional groups positioned ortho or para to the halide leaving group, the SNAr mechanism is favored.

SNAr reaction mechanism

The following is the reaction mechanism of a nucleophilic aromatic substitution of 2,4-dinitrochlorobenzene (1) in a basic solution in water.

Nucleophilic aromatic substitution Nucleophilic aromatic substitution 2,4-dinitrochlorobenzene.svg
Nucleophilic aromatic substitution

Since the nitro group is an activator toward nucleophilic substitution, and a meta director, it is able to stabilize the additional electron density (via resonance) when the aromatic compound is attacked by the hydroxide nucleophile. The resulting intermediate, named the Meisenheimer complex (2a), the ipso carbon is temporarily bonded to the hydroxyl group. This Meisenheimer complex is extra stabilized by the additional electron-withdrawing nitro group (2b).

In order to return to a lower energy state, either the hydroxyl group leaves, or the chloride leaves. In solution, both processes happen. A small percentage of the intermediate loses the chloride to become the product (2,4-dinitrophenol, 3), while the rest return to the reactant (1). Since 2,4-dinitrophenol is in a lower energy state, it will not return to form the reactant, so after some time has passed, the reaction reaches chemical equilibrium that favors the 2,4-dinitrophenol, which is then deprotonated by the basic solution (4).

The formation of the resonance-stabilized Meisenheimer complex is slow because the loss of aromaticity due to nucleophilic attack results in a higher-energy state. By the same coin, the loss of the chloride or hydroxide is fast, because the ring regains aromaticity. Recent work indicates that, sometimes, the Meisenheimer complex is not always a true intermediate but may be the transition state of a 'frontside SN2' process, particularly if stabilization by electron-withdrawing groups is not very strong. [2] A 2019 review argues that such 'concerted SNAr' reactions are more prevalent than previously assumed. [3]

Aryl halides cannot undergo the classic 'backside' SN2 reaction. The carbon-halogen bond is in the plane of the ring because the carbon atom has a trigonal planar geometry. Backside attack is blocked and this reaction is therefore not possible. [4] An SN1 reaction is possible but very unfavourable. It would involve the unaided loss of the leaving group and the formation of an aryl cation. [4] The nitro group is the most commonly encountered activating group, other groups are the cyano and the acyl group. [5] The leaving group can be a halogen or a sulfide. With increasing electronegativity the reaction rate for nucleophilic attack increases. [5] This is because the rate-determining step for an SNAr reaction is attack of the nucleophile and the subsequent breaking of the aromatic system; the faster process is the favourable reforming of the aromatic system after loss of the leaving group. As such, the following pattern is seen with regard to halogen leaving group ability for SNAr: F > Cl ≈ Br > I (i.e. an inverted order to that expected for an SN2 reaction). If looked at from the point of view of an SN2 reaction this would seem counterintuitive, since the C-F bond is amongst the strongest in organic chemistry, when indeed the fluoride is the ideal leaving group for an SNAr due to the extreme polarity of the C-F bond. Nucleophiles can be amines, alkoxides, sulfides and stabilized carbanions. [5]

Nucleophilic aromatic substitution reactions

Some typical substitution reactions on arenes are listed below.

Nucleophilic aromatic substitution is not limited to arenes, however; the reaction takes place even more readily with heteroarenes. Pyridines are especially reactive when substituted in the aromatic ortho position or aromatic para position because then the negative charge is effectively delocalized at the nitrogen position. One classic reaction is the Chichibabin reaction (Aleksei Chichibabin, 1914) in which pyridine is reacted with an alkali-metal amide such as sodium amide to form 2-aminopyridine. [6]

In the compound methyl 3-nitropyridine-4-carboxylate, the meta nitro group is actually displaced by fluorine with cesium fluoride in DMSO at 120 °C. [7]

Nucleophilic aromatic substitution at pyridine NuArSubPyr.png
Nucleophilic aromatic substitution at pyridine

Asymmetric nucleophilic aromatic substitution

With carbon nucleophiles such as 1,3-dicarbonyl compounds the reaction has been demonstrated as a method for the asymmetric synthesis of chiral molecules. [8] First reported in 2005, the organocatalyst (in a dual role with that of a phase transfer catalyst) is derived from cinchonidine (benzylated at N and O).

AsymmetricNucleophilicAromaticSubstitution.png

See also

Related Research Articles

<span class="mw-page-title-main">Aromatic compound</span> Compound containing rings with delocalized pi electrons

Aromatic compounds usually refers to organic compounds "with a chemistry typified by benzene" and "cyclically conjugated " The word "aromatic" originates from the past grouping of molecules based on odor, before their general chemical properties were understood. The current definition of aromatic compounds does not have any relation to their odor. Aromatic compounds are now defined as cyclic compounds satisfying Hückel's Rule. Aromatic compounds have the following general properties:

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.

The following outline is provided as an overview of and topical guide to organic chemistry:

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 via a backside attack, all while the leaving group detaches from the reaction center in a concerted fashion.

In chemistry, an electrophile is a chemical species that forms bonds with nucleophiles by accepting an electron pair. Because electrophiles accept electrons, they are Lewis acids. 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.

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.

In organic chemistry, an aryl halide is an aromatic compound in which one or more hydrogen atoms, directly bonded to an aromatic ring are replaced by a halide. The haloarene are different from haloalkanes because they exhibit many differences in methods of preparation and properties. The most important members are the aryl chlorides, but the class of compounds is so broad that there are many derivatives and applications.

An allylic rearrangement or allylic shift is an organic reaction in which the double bond in an allyl chemical compound shifts to the next carbon atom. It is encountered in nucleophilic substitution.

<span class="mw-page-title-main">Michaelis–Arbuzov reaction</span>

The Michaelis–Arbuzov reaction is the chemical reaction of a trivalent phosphorus ester with an alkyl halide to form a pentavalent phosphorus species and another alkyl halide. The picture below shows the most common types of substrates undergoing the Arbuzov reaction; phosphite esters (1) react to form phosphonates (2), phosphonites (3) react to form phosphinates (4) and phosphinites (5) react to form phosphine oxides (6).

Nucleophilic acyl substitution describes a class of substitution reactions involving nucleophiles and acyl compounds. In this type of reaction, a nucleophile – such as an alcohol, amine, or enolate – displaces the leaving group of an acyl derivative – such as an acid halide, anhydride, or ester. The resulting product is a carbonyl-containing compound in which the nucleophile has taken the place of the leaving group present in the original acyl derivative. Because acyl derivatives react with a wide variety of nucleophiles, and because the product can depend on the particular type of acyl derivative and nucleophile involved, nucleophilic acyl substitution reactions can be used to synthesize a variety of different products.

In organic chemistry, the Hammett equation 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 his 1935 publication.

Radical-nucleophilic aromatic substitution or SRN1 in organic chemistry is a type of substitution reaction in which a certain substituent on an aromatic compound is replaced by a nucleophile through an intermediary free radical species:

<span class="mw-page-title-main">Vicarious nucleophilic substitution</span>

In organic chemistry, the vicarious nucleophilic substitution is a special type of nucleophilic aromatic substitution in which a nucleophile replaces a hydrogen atom on the aromatic ring and not leaving groups such as halogen substituents which are ordinarily encountered in SNAr. This reaction type was reviewed in 1987 by Polish chemists Mieczysław Mąkosza and Jerzy Winiarski.

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">2-Chlorobutane</span> Chemical compound

2-Chlorobutane is a compound with formula C4H9Cl. It is also called sec-butyl chloride. It is a colorless, volatile liquid at room temperature that is not miscible in water.

Ether cleavage refers to chemical substitution reactions that lead to the cleavage of ethers. Due to the high chemical stability of ethers, the cleavage of the C-O bond is uncommon in the absence of specialized reagents or under extreme conditions.

References

  1. Clayden, Jonathan; Greeves, Nick; Warren, Stuart (2012-03-15). Organic Chemistry (Second ed.). Oxford, New York: Oxford University Press. pp. 514–515. ISBN   978-0-19-927029-3.
  2. Neumann CN, Hooker JM, Ritter T (June 2016). "Concerted nucleophilic aromatic substitution with (19)F(-) and (18)F(-)". Nature. 534 (7607): 369–73. doi:10.1038/nature17667. PMC   4911285 . PMID   27281221.
  3. Rohrbach S, Smith AJ, Pang JH, Poole DL, Tuttle T, Chiba S, Murphy JA (November 2019). "Concerted Nucleophilic Aromatic Substitution Reactions". Angewandte Chemie. 58 (46): 16368–16388. doi:10.1002/anie.201902216. PMC   6899550 . PMID   30990931.
  4. 1 2 Clayden J. Organic Chemistry. Oxford University Press.
  5. 1 2 3 Goldstein SW, Bill A, Dhuguru J, Ghoneim O (September 2017). "Nucleophilic Aromatic Substitution Addition and Identification of an Amine". Journal of Chemical Education. 94 (9): 1388–90. Bibcode:2017JChEd..94.1388G. doi:10.1021/acs.jchemed.6b00680.
  6. March J (1966). Advanced Organic Chemistry, Reactions, Mechanisms and Structure (3rd ed.). ISBN   0-471-85472-7.
  7. Tjosaas F, Fiksdahl A (February 2006). "A simple synthetic route to methyl 3-fluoropyridine-4-carboxylate by nucleophilic aromatic substitution". Molecules (Basel, Switzerland). 11 (2): 130–3. doi: 10.3390/11020130 . PMC   6148553 . PMID   17962783.
  8. Bella M, Kobbelgaard S, Jørgensen KA (March 2005). "Organocatalytic regio- and asymmetric C-selective S(N)Ar reactions-stereoselective synthesis of optically active spiro-pyrrolidone-3,3'-oxoindoles". Journal of the American Chemical Society. 127 (11): 3670–1. doi:10.1021/ja050200g. PMID   15771481.