fbpx
Wikipedia

Nucleophilic substitution

In chemistry, a nucleophilic substitution is a class of chemical reactions in which an electron-rich chemical species (known as a nucleophile) replaces a functional group within another electron-deficient molecule (known as the electrophile). The molecule that contains the electrophile and the leaving functional group is called the substrate.[1][2]

The most general form of the reaction may be given as the following:

The electron pair (:) from the nucleophile (Nuc) attacks the substrate (R−LG) and bonds with it. Simultaneously, the leaving group (LG) departs with an electron pair. The principal product in this case is R−Nuc. The nucleophile may be electrically neutral or negatively charged, whereas the substrate is typically neutral or positively charged.

An example of nucleophilic substitution is the hydrolysis of an alkyl bromide, R-Br under basic conditions, where the attacking nucleophile is hydroxyl (OH) and the leaving group is bromide (Br).

Nucleophilic substitution reactions are common in organic chemistry. Nucleophiles often attack a saturated aliphatic carbon. Less often, they may attack an aromatic or unsaturated carbon.[3]

Saturated carbon centres Edit

SN1 and SN2 reactions Edit

 
A graph showing the relative reactivities of the different alkyl halides towards SN1 and SN2 reactions (also see Table 1).

In 1935, Edward D. Hughes and Sir Christopher Ingold studied nucleophilic substitution reactions of alkyl halides and related compounds. They proposed that there were two main mechanisms at work, both of them competing with each other. The two main mechanisms were the SN1 reaction and the SN2 reaction, where S stands for substitution, N stands for nucleophilic, and the number represents the kinetic order of the reaction.[4]

In the SN2 reaction, the addition of the nucleophile and the elimination of leaving group take place simultaneously (i.e. a concerted reaction). SN2 occurs when the central carbon atom is easily accessible to the nucleophile.[5]

Nucleophilic substitution at carbon
 
 
SN2 mechanism

In SN2 reactions, there are a few conditions that affect the rate of the reaction. First of all, the 2 in SN2 implies that there are two concentrations of substances that affect the rate of reaction: substrate (Sub) and nucleophile. The rate equation for this reaction would be Rate=k[Sub][Nuc]. For a SN2 reaction, an aprotic solvent is best, such as acetone, DMF, or DMSO. Aprotic solvents do not add protons (H+ ions) into solution; if protons were present in SN2 reactions, they would react with the nucleophile and severely limit the reaction rate. Since this reaction occurs in one step, steric effects drive the reaction speed. In the intermediate step, the nucleophile is 185 degrees from the leaving group and the stereochemistry is inverted as the nucleophile bonds to make the product. Also, because the intermediate is partially bonded to the nucleophile and leaving group, there is no time for the substrate to rearrange itself: the nucleophile will bond to the same carbon that the leaving group was attached to. A final factor that affects reaction rate is nucleophilicity; the nucleophile must attack an atom other than a hydrogen.

By contrast the SN1 reaction involves two steps. SN1 reactions tend to be important when the central carbon atom of the substrate is surrounded by bulky groups, both because such groups interfere sterically with the SN2 reaction (discussed above) and because a highly substituted carbon forms a stable carbocation.

Nucleophilic substitution at carbon
 
SN1 mechanism

Like SN2 reactions, there are quite a few factors that affect the reaction rate of SN1 reactions. Instead of having two concentrations that affect the reaction rate, there is only one, substrate. The rate equation for this would be Rate=k[Sub]. Since the rate of a reaction is only determined by its slowest step, the rate at which the leaving group "leaves" determines the speed of the reaction. This means that the better the leaving group, the faster the reaction rate. A general rule for what makes a good leaving group is the weaker the conjugate base, the better the leaving group. In this case, halogens are going to be the best leaving groups, while compounds such as amines, hydrogen, and alkanes are going to be quite poor leaving groups. As SN2 reactions were affected by sterics, SN1 reactions are determined by bulky groups attached to the carbocation. Since there is an intermediate that actually contains a positive charge, bulky groups attached are going to help stabilize the charge on the carbocation through resonance and distribution of charge. In this case, tertiary carbocation will react faster than a secondary which will react much faster than a primary. It is also due to this carbocation intermediate that the product does not have to have inversion. The nucleophile can attack from the top or the bottom and therefore create a racemic product. It is important to use a protic solvent, water and alcohols, since an aprotic solvent could attack the intermediate and cause unwanted product. It does not matter if the hydrogens from the protic solvent react with the nucleophile since the nucleophile is not involved in the rate determining step.

Table 1. Nucleophilic substitutions on RX (an alkyl halide or equivalent)
Factor SN1 SN2 Comments
Kinetics Rate = k[RX] Rate = k[RX][Nuc]
Primary alkyl Never unless additional stabilising groups present Good unless a hindered nucleophile is used
Secondary alkyl Moderate Moderate
Tertiary alkyl Excellent Never Elimination likely if heated or if strong base used
Leaving group Important Important For halogens,
I > Br > Cl >> F
Nucleophilicity Unimportant Important
Preferred solvent Polar protic Polar aprotic
Stereochemistry Racemisation (+ partial inversion possible) Inversion
Rearrangements Common Rare Side reaction
Eliminations Common, especially with basic nucleophiles Only with heat & basic nucleophiles Side reaction
esp. if heated

Reactions Edit

There are many reactions in organic chemistry involving this type of mechanism. Common examples include:

R−XR−H using LiAlH4   (SN2)
R−Br + OHR−OH + Br (SN2) or
R−Br + H2O → R−OH + HBr   (SN1)
R−Br + OR'R−OR' + Br   (SN2)

Borderline mechanism Edit

An example of a substitution reaction taking place by a so-called borderline mechanism as originally studied by Hughes and Ingold[6] is the reaction of 1-phenylethyl chloride with sodium methoxide in methanol.

 

The reaction rate is found to the sum of SN1 and SN2 components with 61% (3,5 M, 70 °C) taking place by the latter.

Other mechanisms Edit

Besides SN1 and SN2, other mechanisms are known, although they are less common. The SNi mechanism is observed in reactions of thionyl chloride with alcohols, and it is similar to SN1 except that the nucleophile is delivered from the same side as the leaving group.

Nucleophilic substitutions can be accompanied by an allylic rearrangement as seen in reactions such as the Ferrier rearrangement. This type of mechanism is called an SN1' or SN2' reaction (depending on the kinetics). With allylic halides or sulphonates, for example, the nucleophile may attack at the γ unsaturated carbon in place of the carbon bearing the leaving group. This may be seen in the reaction of 1-chloro-2-butene with sodium hydroxide to give a mixture of 2-buten-1-ol and 1-buten-3-ol:

 

The Sn1CB mechanism appears in inorganic chemistry. Competing mechanisms exist.[7][8]

In organometallic chemistry the nucleophilic abstraction reaction occurs with a nucleophilic substitution mechanism.

Unsaturated carbon centres Edit

Nucleophilic substitution via the SN1 or SN2 mechanism does not generally occur with vinyl or aryl halides or related compounds. Under certain conditions nucleophilic substitutions may occur, via other mechanisms such as those described in the nucleophilic aromatic substitution article.

When the substitution occurs at the carbonyl group, the acyl group may undergo nucleophilic acyl substitution. This is the normal mode of substitution with carboxylic acid derivatives such as acyl chlorides, esters and amides.

References Edit

  1. ^ March, J. (1992). Advanced Organic Chemistry (4th ed.). New York: Wiley. ISBN 9780471601807.
  2. ^ R. A. Rossi, R. H. de Rossi, Aromatic Substitution by the SRN1 Mechanism, ACS Monograph Series No. 178, American Chemical Society, 1983. ISBN 0-8412-0648-1.
  3. ^ L. G. Wade, Organic Chemistry, 5th ed., Prentice Hall, Upper Saddle River, New Jersey, 2003.
  4. ^ S. R. Hartshorn, Aliphatic Nucleophilic Substitution, Cambridge University Press, London, 1973. ISBN 0-521-09801-7
  5. ^ Introducing Aliphatic Substitution with a Discovery Experiment Using Competing Electrophiles Timothy P. Curran, Amelia J. Mostovoy, Margaret E. Curran, and Clara Berger Journal of Chemical Education 2016 93 (4), 757-761 doi:10.1021/acs.jchemed.5b00394
  6. ^ 253. Reaction kinetics and the Walden inversion. Part II. Homogeneous hydrolysis, alcoholysis, and ammonolysis of -phenylethyl halides Edward D. Hughes, Christopher K. Ingold and Alan D. Scott, J. Chem. Soc., 1937, 1201 doi:10.1039/JR9370001201
  7. ^ N.S.Imyanitov. Electrophilic Bimolecular Substitution as an Alternative to Nucleophilic Monomolecular Substitution in Inorganic and Organic Chemistry. J. Gen. Chem. USSR (Engl. Transl.) 1990; 60 (3); 417-419.
  8. ^ Unimolecular Nucleophilic Substitution does not Exist! / N.S.Imyanitov. SciTecLibrary

External links Edit

nucleophilic, substitution, chemistry, nucleophilic, substitution, class, chemical, reactions, which, electron, rich, chemical, species, known, nucleophile, replaces, functional, group, within, another, electron, deficient, molecule, known, electrophile, molec. In chemistry a nucleophilic substitution is a class of chemical reactions in which an electron rich chemical species known as a nucleophile replaces a functional group within another electron deficient molecule known as the electrophile The molecule that contains the electrophile and the leaving functional group is called the substrate 1 2 The most general form of the reaction may be given as the following Nuc R LG R Nuc LG displaystyle text Nuc mathbf ce R LG gt R Nuc text LG mathbf The electron pair from the nucleophile Nuc attacks the substrate R LG and bonds with it Simultaneously the leaving group LG departs with an electron pair The principal product in this case is R Nuc The nucleophile may be electrically neutral or negatively charged whereas the substrate is typically neutral or positively charged An example of nucleophilic substitution is the hydrolysis of an alkyl bromide R Br under basic conditions where the attacking nucleophile is hydroxyl OH and the leaving group is bromide Br R Br OH R OH Br displaystyle ce R Br OH gt R OH Br Nucleophilic substitution reactions are common in organic chemistry Nucleophiles often attack a saturated aliphatic carbon Less often they may attack an aromatic or unsaturated carbon 3 Contents 1 Saturated carbon centres 1 1 SN1 and SN2 reactions 2 Reactions 2 1 Borderline mechanism 2 2 Other mechanisms 3 Unsaturated carbon centres 4 References 5 External linksSaturated carbon centres EditSN1 and SN2 reactions Edit nbsp A graph showing the relative reactivities of the different alkyl halides towards SN1 and SN2 reactions also see Table 1 In 1935 Edward D Hughes and Sir Christopher Ingold studied nucleophilic substitution reactions of alkyl halides and related compounds They proposed that there were two main mechanisms at work both of them competing with each other The two main mechanisms were the SN1 reaction and the SN2 reaction where S stands for substitution N stands for nucleophilic and the number represents the kinetic order of the reaction 4 In the SN2 reaction the addition of the nucleophile and the elimination of leaving group take place simultaneously i e a concerted reaction SN2 occurs when the central carbon atom is easily accessible to the nucleophile 5 Nucleophilic substitution at carbon nbsp nbsp SN2 mechanismIn SN2 reactions there are a few conditions that affect the rate of the reaction First of all the 2 in SN2 implies that there are two concentrations of substances that affect the rate of reaction substrate Sub and nucleophile The rate equation for this reaction would be Rate k Sub Nuc For a SN2 reaction an aprotic solvent is best such as acetone DMF or DMSO Aprotic solvents do not add protons H ions into solution if protons were present in SN2 reactions they would react with the nucleophile and severely limit the reaction rate Since this reaction occurs in one step steric effects drive the reaction speed In the intermediate step the nucleophile is 185 degrees from the leaving group and the stereochemistry is inverted as the nucleophile bonds to make the product Also because the intermediate is partially bonded to the nucleophile and leaving group there is no time for the substrate to rearrange itself the nucleophile will bond to the same carbon that the leaving group was attached to A final factor that affects reaction rate is nucleophilicity the nucleophile must attack an atom other than a hydrogen By contrast the SN1 reaction involves two steps SN1 reactions tend to be important when the central carbon atom of the substrate is surrounded by bulky groups both because such groups interfere sterically with the SN2 reaction discussed above and because a highly substituted carbon forms a stable carbocation Nucleophilic substitution at carbon nbsp SN1 mechanismLike SN2 reactions there are quite a few factors that affect the reaction rate of SN1 reactions Instead of having two concentrations that affect the reaction rate there is only one substrate The rate equation for this would be Rate k Sub Since the rate of a reaction is only determined by its slowest step the rate at which the leaving group leaves determines the speed of the reaction This means that the better the leaving group the faster the reaction rate A general rule for what makes a good leaving group is the weaker the conjugate base the better the leaving group In this case halogens are going to be the best leaving groups while compounds such as amines hydrogen and alkanes are going to be quite poor leaving groups As SN2 reactions were affected by sterics SN1 reactions are determined by bulky groups attached to the carbocation Since there is an intermediate that actually contains a positive charge bulky groups attached are going to help stabilize the charge on the carbocation through resonance and distribution of charge In this case tertiary carbocation will react faster than a secondary which will react much faster than a primary It is also due to this carbocation intermediate that the product does not have to have inversion The nucleophile can attack from the top or the bottom and therefore create a racemic product It is important to use a protic solvent water and alcohols since an aprotic solvent could attack the intermediate and cause unwanted product It does not matter if the hydrogens from the protic solvent react with the nucleophile since the nucleophile is not involved in the rate determining step Table 1 Nucleophilic substitutions on RX an alkyl halide or equivalent Factor SN1 SN2 CommentsKinetics Rate k RX Rate k RX Nuc Primary alkyl Never unless additional stabilising groups present Good unless a hindered nucleophile is usedSecondary alkyl Moderate ModerateTertiary alkyl Excellent Never Elimination likely if heated or if strong base usedLeaving group Important Important For halogens I gt Br gt Cl gt gt FNucleophilicity Unimportant ImportantPreferred solvent Polar protic Polar aproticStereochemistry Racemisation partial inversion possible InversionRearrangements Common Rare Side reactionEliminations Common especially with basic nucleophiles Only with heat amp basic nucleophiles Side reaction esp if heatedReactions EditThere are many reactions in organic chemistry involving this type of mechanism Common examples include Organic reductions with hydrides for exampleR X R H using LiAlH4 SN2 dd Hydrolysis reactions such asR Br OH R OH Br SN2 or R Br H2O R OH HBr SN1 dd Williamson ether synthesisR Br OR R OR Br SN2 dd The Wenker synthesis a ring closing reaction of aminoalcohols The Finkelstein reaction a halide exchange reaction Phosphorus nucleophiles appear in the Perkow reaction and the Michaelis Arbuzov reaction The Kolbe nitrile synthesis the reaction of alkyl halides with cyanides Borderline mechanism Edit An example of a substitution reaction taking place by a so called borderline mechanism as originally studied by Hughes and Ingold 6 is the reaction of 1 phenylethyl chloride with sodium methoxide in methanol nbsp The reaction rate is found to the sum of SN1 and SN2 components with 61 3 5 M 70 C taking place by the latter Other mechanisms Edit Besides SN1 and SN2 other mechanisms are known although they are less common The SNi mechanism is observed in reactions of thionyl chloride with alcohols and it is similar to SN1 except that the nucleophile is delivered from the same side as the leaving group Nucleophilic substitutions can be accompanied by an allylic rearrangement as seen in reactions such as the Ferrier rearrangement This type of mechanism is called an SN1 or SN2 reaction depending on the kinetics With allylic halides or sulphonates for example the nucleophile may attack at the g unsaturated carbon in place of the carbon bearing the leaving group This may be seen in the reaction of 1 chloro 2 butene with sodium hydroxide to give a mixture of 2 buten 1 ol and 1 buten 3 ol CH 3 CH CH CH 2 Cl CH 3 CH CH CH 2 OH CH 3 CH OH CH CH 2 displaystyle ce CH3CH CH CH2 Cl gt CH3CH CH CH2 OH CH3CH OH CH CH2 nbsp The Sn1CB mechanism appears in inorganic chemistry Competing mechanisms exist 7 8 In organometallic chemistry the nucleophilic abstraction reaction occurs with a nucleophilic substitution mechanism Unsaturated carbon centres EditNucleophilic substitution via the SN1 or SN2 mechanism does not generally occur with vinyl or aryl halides or related compounds Under certain conditions nucleophilic substitutions may occur via other mechanisms such as those described in the nucleophilic aromatic substitution article When the substitution occurs at the carbonyl group the acyl group may undergo nucleophilic acyl substitution This is the normal mode of substitution with carboxylic acid derivatives such as acyl chlorides esters and amides References Edit March J 1992 Advanced Organic Chemistry 4th ed New York Wiley ISBN 9780471601807 R A Rossi R H de Rossi Aromatic Substitution by the SRN1 Mechanism ACS Monograph Series No 178 American Chemical Society 1983 ISBN 0 8412 0648 1 L G Wade Organic Chemistry 5th ed Prentice Hall Upper Saddle River New Jersey 2003 S R Hartshorn Aliphatic Nucleophilic Substitution Cambridge University Press London 1973 ISBN 0 521 09801 7 Introducing Aliphatic Substitution with a Discovery Experiment Using Competing Electrophiles Timothy P Curran Amelia J Mostovoy Margaret E Curran and Clara Berger Journal of Chemical Education 2016 93 4 757 761 doi 10 1021 acs jchemed 5b00394 253 Reaction kinetics and the Walden inversion Part II Homogeneous hydrolysis alcoholysis and ammonolysis of phenylethyl halides Edward D Hughes Christopher K Ingold and Alan D Scott J Chem Soc 1937 1201 doi 10 1039 JR9370001201 N S Imyanitov Electrophilic Bimolecular Substitution as an Alternative to Nucleophilic Monomolecular Substitution in Inorganic and Organic Chemistry J Gen Chem USSR Engl Transl 1990 60 3 417 419 Unimolecular Nucleophilic Substitution does not Exist N S Imyanitov SciTecLibraryExternal links Edit nbsp Wikiquote has quotations related to Nucleophilic substitution Retrieved from https en wikipedia org w index php title Nucleophilic substitution amp oldid 1175220065, wikipedia, wiki, book, books, library,

article

, read, download, free, free download, mp3, video, mp4, 3gp, jpg, jpeg, gif, png, picture, music, song, movie, book, game, games.