Sei sulla pagina 1di 301

FREE RADICAL MECHANISMS

A. Free radical substitution


1. What is free radical substitution? . . . A brief explanation of the terms free radical and substitution. 2. methane and chlorine . . . Free radical substitution of hydrogen atoms in methane by chlorine atoms. Includes multiple substitution if you need it. 3. methane and bromine . . . Free radical substitution of hydrogen atoms in methane by bromine atoms. Includes multiple substitution if you need it. 4. methylbenzene and chlorine . . . Free radical substitution of hydrogen atoms in the methyl group in methylbenzene by chlorine atoms in the presence of UV light (sunlight). Includes multiple substitution if you need it.

B. Free radical addition


1. polymerisation of ethene . . . The mechanism for the polymerisation of ethene under free radical conditions. 2. alkenes with hydrogen bromide . . . The mechanism for the addition of hydrogen bromide to alkenes under free radical conditions. This is often known as the "peroxide effect".

FREE RADICAL SUBSTITUTION


Substitution reactions
These are reactions in which one atom in a molecule is replaced by another atom or group of atoms. Free radical substitution for A' level purposes involves breaking a carbon-hydrogen bond in alkanes such as methane ethane propane CH4 CH3CH3 CH3CH2CH3

A new bond is then formed to something else. It also happens in alkyl groups like methyl, ethyl (and so on) wherever these appear in more complicated molecules. methyl ethyl CH3 CH3CH2

For example, ethanoic acid is CH3COOH and contains a methyl group. The carbon-hydrogen bonds in the methyl group behave just like those in methane, and can be broken and replaced by something else in the same way. A simple example of substitution is the reaction between methane and chlorine in the presence of UV light (or sunlight). CH4 + Cl2 CH3Cl + HCl

Notice that one of the hydrogen atoms in the methane has been replaced by a chlorine atom. That's substitution.

Free radical reactions


Free radicals are atoms or groups of atoms which have a single unpaired electron. A free radical substitution reaction is one involving these radicals. Free radicals are formed if a bond splits evenly - each atom getting one of the two electrons. The name given to this is

homolytic fission.
Note: If a bond were to split unevenly (one atom getting both electrons, and the other none), ions would be formed. The atom that got both electrons would become negatively charged, while the other one would become positive. This is called heterolytic fission. Warning! It is important that you get these terms the right way round. "Fission" is obvious - it just means "splitting". "Homo" and "hetero" are used in the sense of "same" (homo) or "different" (hetero). This is just like their use in "homosexual" or "heterosexual". So, homolytic fission is splitting a bond to produce two particles which are the same in the sense that they both have a single unpaired electron (both are free radicals). Heterolytic fission produces two particles which are different because one is a positive ion and the other a negative ion.

To show that a species (either an atom or a group of atoms) is a free radical, the symbol is written with a dot attached to show the unpaired electron. For example: a chlorine radical a methyl radical Cl CH3

Note: If you wanted to be fussy, the dot showing the electron really ought to be written next to the carbon atom in the methyl radical, because that's the atom with the unpaired electron - in other words as CH3. This isn't normally done unless the radical gets more complicated. Examples of this will crop up when you look at the mechanisms.

THE REACTION BETWEEN METHANE AND CHLORINE


A Free Radical Substitution Reaction

This page gives you the facts and a simple, uncluttered mechanism for the free radical substitution reaction between methane and chlorine. If you want the mechanism explained to you in detail, there is a link at the bottom of the page. The facts If a mixture of methane and chlorine is exposed to a flame, it explodes - producing carbon and hydrogen chloride. This isn't a very useful reaction! The reaction we are going to explore is a more gentle one between methane and chlorine in the presence of ultraviolet light - typically sunlight. This is a good example of a photochemical reaction - a reaction brought about by light.
Note: These reactions are sometimes described as examples of photocatalysis - reactions catalysed by light. It is better to use the term "photochemical" and keep the keep the word "catalysis" for reactions speeded up by actual substances rather than light.

CH4 + Cl2

CH3Cl + HCl

The organic product is chloromethane. One of the hydrogen atoms in the methane has been replaced by a chlorine atom, so this is a substitution reaction. However, the reaction doesn't stop there, and all the hydrogens in the methane can in turn be replaced by chlorine atoms. Multiple substitution is dealt with on a separate page, and you will find a link to that at the bottom of this page. The mechanism The mechanism involves a chain reaction. During a chain reaction, for every reactive species you start off with, a new one is generated at the end - and this keeps the process going.
Species: a useful word which is used in chemistry to mean any sort of particle you want it to mean. It covers molecules, ions, atoms, or (in this case) free radicals.

The over-all process is known as free radical substitution, or as a free radical chain reaction.
Note: If you aren't sure about the words free radical or substitution, read the page What is free radical substitution? Use the BACK button on your browser to return quickly to this page.

Chain initiation The chain is initiated (started) by UV light breaking a chlorine molecule into free radicals. Cl2 2Cl

Chain propagation reactions These are the reactions which keep the chain going. CH4 + Cl CH3 + Cl2 CH3 + HCl

CH3Cl + Cl

Chain termination reactions These are reactions which remove free radicals from the system without replacing them by new ones. 2Cl CH3 CH3 + Cl + CH3 Cl2 CH3Cl CH3CH3

EXPLAINING THE REACTION BETWEEN METHANE AND CHLORINE

A Free Radical Substitution Reaction This page guides you through the mechanism for the substitution of one of the hydrogen atoms in methane by one chlorine atom. Multiple substitution is covered separately, and you will find a link at the bottom of the

page. We are going to talk through this mechanism in a very detailed way so that you get a feel for what is going on. You couldn't possibly do the same thing in an exam. At the bottom of the page, you will find the condensed down version corresponding to the sort of answer you would produce in an exam. The role of the UV light The ultraviolet light is simply a source of energy, and is being used to break bonds. In fact, the energies in UV are exactly right to break the bonds in chlorine molecules to produce chlorine atoms.

Note: Only the outer electrons of the chlorine are shown. Notice also that it is quite acceptable to use a simple view of atomic structure. There is no point in using a complicated model of the atom if a simple one will do the job.

Because we want to stress the fact that the chlorine atoms have single unpaired electrons, then we call them chlorine free radicals - or more usually just chlorine radicals. To show that a species (either an atom or a group of atoms) is a free radical, the symbol is written with a dot attached to show the unpaired electron. The splitting of the chlorine molecule would be shown as: Cl2 2Cl

Free radicals are formed if a bond splits evenly - each atom getting one of the two electrons. The name given to this is homolytic fission.

What happens to the chlorine radicals? There's nothing magic about reaction mechanisms. Reactions happen because things hit each other. If the conditions are right something useful might happen. In this case, you need to think about what the chlorine radicals are likely to hit, and what could happen as a result of that collision. At the moment the mixture contains

lots of methane molecules lots of chlorine molecules (only a few will have been fractured by the UV light) a few chlorine radicals

Let's start with the unproductive collisions. The least likely collision is between two chlorine radicals. There aren't very many of them in the mixture and so the chances of them hitting each other are relatively small. If they do collide, they will combine to form a chlorine molecule. That's worse than useless because it removes the active free radicals from the system. 2Cl Cl2

A chlorine radical could also hit a chlorine molecule. If this happens there could possibly be an exchange of chlorine atoms, but nothing new would be formed. It is just a wasted collision. Cl + Cl-Cl Cl-Cl + Cl
Note: There is no difference between the chlorine atoms shown in bold type or ordinary type. They are shown differently so that the exchange is made clear.

The productive collision happens if a chlorine radical hits a methane molecule.

The chlorine radical removes a hydrogen atom from the methane. That hydrogen atom only needs to bring one electron with it to form a new bond to the chlorine, and so one electron is left behind on the carbon atom. A new free radical is formed this time a methyl radical, CH3 . CH4 + Cl CH3 + HCl

What happens to the methyl radicals? It depends what they collide with. There are three interesting collisions which need to be explored. Two of these involve a setback to the reaction, and only one is useful. Luckily, the two unhelpful collisions don't happen very often, because they involve collisions between two free radicals - and there won't be many of these present in the mixture at any one time. CH3 CH3 + Cl + CH3 CH3Cl CH3CH3

Even though the first reaction seems to produce what you want, the problem with both of these reactions is that they use up the free radicals in the system - we'll come back to that problem shortly. The second reaction, of course, also introduces an impurity into the mixture. So what is the useful collision? If a methyl radical hits a chlorine molecule (something that's quite likely to occur), the following change can happen: CH3 + Cl2 CH3Cl + Cl

The methyl radical takes one of the chlorine atoms to form chloromethane (which is what we want to make), but in the

process generates another chlorine radical. This new chlorine radical can now go through the whole sequence again, and at the end will produce yet another chlorine radical - and so on and so on. The process is described as a free radical chain reaction. The chain continues because for every chlorine radical that goes in at the beginning, a new one is generated at the end. Chain termination Does this mean that one tiny burst of UV light, splitting one chlorine molecule into two free radicals, is enough to convert a whole reactions-worth of methane and chlorine into chloromethane and HCl? Sadly, no! As we've seen, there are collisions which result in the removal of free radicals without producing any new ones. These radicals can only be replaced by starting the process all over again with a new burst of light energy. In practice, then, the chains propagate many thousands of times, but eventually any chain will be brought to an end by one of these chain termination processes.

Simplifying all this for exam purposes: The over-all process is known as free radical substitution, or as a free radical chain reaction. Chain initiation The chain is initiated (started) by UV light breaking a chlorine molecule into free radicals. Cl2 2Cl

Chain propagation reactions These are the reactions which keep the chain going. CH4 + Cl CH3 + Cl2 CH3 + HCl

CH3Cl + Cl

Chain termination reactions

These are reactions which remove free radicals from the system without replacing them by new ones. 2Cl CH3 CH3 + Cl + CH3 Cl2 CH3Cl CH3CH3

MULTIPLE SUBSTITUTION IN THE METHANE AND CHLORINE REACTION


Warning! We are just about to muddy the water quite considerably! Don't go on until you are sure that you understand the mechanism for the production of chloromethane - and are confident that you could write it in an exam. If you aren't sure about it, go back to that reaction and look at it again. It would be worth checking your syllabus and past exam papers to see if you need to know about these further substitution reactions.

The facts
When a mixture of methane and chlorine is exposed to ultraviolet light - typically sunlight - a substitution reaction occurs and the organic product is chloromethane. CH4 + Cl2 CH3Cl + HCl

However, the reaction doesn't stop there, and all the hydrogens in the methane can in turn be replaced by chlorine atoms. That means that you could get any of chloromethane, dichloromethane, trichloromethane or tetrachloromethane. CH4 + Cl2 CH3Cl + Cl2 CH2Cl2 + Cl2 CHCl3 + Cl2 CH3Cl + CH2Cl2 CHCl3 CCl4 + HCl + HCl + HCl HCl

You might think that you could control which product you got by the proportions of methane and chlorine you used, but it isn't as simple as that. If you use enough chlorine you will eventually get CCl4, but any other proportions will always lead to a mixture of products.

The mechanisms
The formation of multiple substitution products like di-, tri- and tetrachloromethane can be explained in just the same sort of way as the formation of the original chloromethane. You just have to look at the likely collisions as the reaction progresses. Making dichloromethane You will remember that the over-all equation for the first stage of the reaction is CH4 + Cl2 CH3Cl + HCl

As the reaction proceeds, the methane is getting used up and chloromethane is taking its place. That means that the argument about what a chlorine radical is likely to hit changes during the course of the reaction. As time goes by there is an increasing chance of it hitting a chloromethane molecule rather than a methane molecule. When that happens, the chlorine radical can take a hydrogen from the chloromethane just as well as it could from a methane. In this new case: CH3Cl + Cl CH2Cl + HCl

Notice: The dot representing the electron has been moved against the carbon which is the atom with the unpaired electron. It would be potentially confusing to leave it next to the chlorine.

The chloromethyl radical formed can then interact with a chlorine molecule in a new propagation step . . . CH2Cl + Cl2 CH2Cl2 + Cl

. . . and so dichloromethane is formed and a chlorine radical regenerated. These propagation steps continue until the chain is terminated by any two radicals colliding and combining together. Making tri- and tetrachloromethane Obviously, as time goes on, there is an increasing chance of the dichloromethane being hit by a chlorine radical - producing these propagation steps giving trichloromethane: CH2Cl2 + Cl CHCl2 + Cl2 CHCl2 + HCl CHCl3 + Cl

Care! Don't just skip lightly over these equations. Look carefully at each one so that you understand what is happening, and can relate it to what has gone before. Talk through the equations with yourself. For example: "A chlorine radical hits the dichloromethane molecule and steals a hydrogen. That leaves a new radical (I don't know what it's called, but that doesn't really matter, as long as I can work out its formula if I have to!), which then bumps into a chlorine molecule - etc, etc." Doing this helps you to focus properly on the equations. If you just read them quickly, you'll have forgotten all about them again in 15 seconds!

As the amount of trichloromethane builds up, then you will get these steps giving tetrachloromethane: CHCl3 + Cl CCl3 + Cl2 CCl3 + HCl CCl4 + Cl

This is why you will always get a mixture of products whatever the reaction proportions of methane and chlorine you use. The whole process is simply governed by chance. Having produced some chloromethane there is no way that you can prevent it from being hit by chlorine radicals, and similarly for

dichloromethane and trichloromethane. Trying to produce mainly one product If you wanted tetrachloromethane, you could of course get it by using a large excess of chlorine, so that eventually all the hydrogens would be replaced. If you wanted mainly chloromethane, you could favour this by using a huge excess of methane so that the chances were always greater of a chlorine radical hitting a methane rather than anything else - but even so, you would still get some mixture of products. There is no obvious way of getting mainly dichloromethane or trichloromethane.

SIDE REACTIONS IN THE METHANE AND CHLORINE REACTION


You may remember that one of the chain termination steps produces ethane, CH 3CH3. CH3 + CH3 CH3CH3

If chlorine radicals hit that, you are going to get chloroethane and dichloroethane and so on - and in the course of those reactions you will get ethyl radicals which could themselves become involved in chain termination steps leading to propane (from methyl radical hitting ethyl radical) or butane (from two ethyl radicals combining), which could then start to undergo substitution - and on and on! To be honest, all of these side products are going to be present in very small amounts because the reaction producing ethane won't, by chance, happen very often, but it nicely illustrates a typical organic chemistry problem - when you do a reaction in the lab to produce an organic chemical, a high proportion of your time is spent in purifying the product from all the side reactions that have gone on!

THE REACTION BETWEEN METHANE AND BROMINE


A Free Radical Substitution Reaction
This page gives you the facts and a simple, uncluttered mechanism for the free radical substitution reaction between methane and bromine. If you want the mechanism explained to you in detail, there is a link at the bottom of the page. The facts This reaction between methane and bromine happens in the presence of ultraviolet light - typically sunlight. This is a good example of a photochemical reaction - a reaction brought about by light.
Note: These reactions are sometimes described as examples of photocatalysis - reactions catalysed by light. It is better to use the term "photochemical" and keep the keep the word "catalysis" for reactions speeded up by actual substances rather than light.

CH4 + Br2

CH3Br + HBr

The organic product is bromomethane. One of the hydrogen atoms in the methane has been replaced by a bromine atom, so this is a substitution reaction. However, the reaction doesn't stop there, and all the hydrogens in the methane can in turn be replaced by bromine atoms. Multiple substitution is dealt with on a separate page, and you will find a link to that at the bottom of this page.
Warning! Check your syllabus at this point. If your syllabus wants you to know about the free radical substitution reaction between methane and chlorine as well as this one, don't waste time trying to learn both

mechanisms. The two mechanisms are identical. You just need to learn one of them. If you are asked for the other one, all you need to do is to write bromine, say, instead of chlorine. In writing the bromine mechanisms on these pages, that's exactly what I've done! If you read both chlorine and bromine versions, you'll find them boringly repetitive!

The mechanism The mechanism involves a chain reaction. During a chain reaction, for every reactive species you start off with, a new one is generated at the end - and this keeps the process going.
Species: a useful word which is used in chemistry to mean any sort of particle you want it to mean. It covers molecules, ions, atoms, or (in this case) free radicals.

The over-all process is known as free radical substitution, or as a free radical chain reaction.
Note: If you aren't sure about the words free radical or substitution, read the page What is free radical substitution? Use the BACK button on your browser to return quickly to this page.

Chain initiation The chain is initiated (started) by UV light breaking a bromine molecule into free radicals. Br2 2Br

Chain propagation reactions These are the reactions which keep the chain going. CH4 + Br CH3 + HBr

CH3

+ Br2

CH3Br + Br

Chain termination reactions These are reactions which remove free radicals from the system without replacing them by new ones. 2Br CH3 CH3 + Br + CH3 Br2 CH3Br CH3CH3

EXPLAINING THE REACTION BETWEEN METHANE AND BROMINE


A Free Radical Substitution Reaction
This page guides you through the mechanism for the substitution of one of the hydrogen atoms in methane by one bromine atom. Multiple substitution is covered separately, and you will find a link at the bottom of the page. We are going to talk through this mechanism in a very detailed way so that you get a feel for what is going on. You couldn't possibly do the same thing in an exam. At the bottom of the page, you will find the condensed down version corresponding to the sort of answer you would produce in an exam. The role of the UV light The ultraviolet light is simply a source of energy, and is being used to break bonds. In fact, the energies in UV are exactly right to break the bonds in bromine molecules to produce bromine atoms.

Note: Only the outer electrons of the bromine are shown. Notice also that it is quite acceptable to use a simple view of atomic structure. There is no point in using a complicated model of the atom if a simple one will do the job.

Because we want to stress the fact that the bromine atoms have single unpaired electrons, then we call them bromine free radicals - or more usually just bromine radicals. To show that a species (either an atom or a group of atoms) is a free radical, the symbol is written with a dot attached to show the unpaired electron. The splitting of the bromine molecule would be shown as: Br2 2Br

Free radicals are formed if a bond splits evenly - each atom getting one of the two electrons. The name given to this is homolytic fission. What happens to the bromine radicals? There's nothing magic about reaction mechanisms. Reactions happen because things hit each other. If the conditions are right something useful might happen. In this case, you need to think about what the bromine radicals are likely to hit, and what could happen as a result of that collision. At the moment the mixture contains

lots of methane molecules lots of bromine molecules (only a few will have been fractured by the UV light) a few bromine radicals

Let's start with the unproductive collisions. The least likely collision is between two bromine radicals. There aren't very many of them in the mixture and so the chances of them hitting each other are relatively small. If they do collide, they will combine to form a bromine molecule. That's worse than useless because it removes the active free radicals from the

system. 2Br Br2

A bromine radical could also hit a bromine molecule. If this happens there could possibly be an exchange of bromine atoms, but nothing new would be formed. It is just a wasted collision. Br + Br-Br Br-Br + Br
Note: There is no difference between the bromine atoms shown in bold type or ordinary type. They are shown differently so that the exchange is made clear.

The productive collision happens if a bromine radical hits a methane molecule.

The bromine radical removes a hydrogen atom from the methane. That hydrogen atom only needs to bring one electron with it to form a new bond to the bromine, and so one electron is left behind on the carbon atom. A new free radical is formed this time a methyl radical, CH3 . CH4 + Br CH3 + HBr

What happens to the methyl radicals? It depends what they collide with. There are three interesting collisions which need to be explored. Two of these involve a setback to the reaction, and only one is useful. Luckily, the two unhelpful collisions don't happen very often, because they involve collisions between two free radicals - and there won't be many of these present in the mixture at any one

time. CH3 CH3 + Br + CH3 CH3Br CH3CH3

Even though the first reaction seems to produce what you want, the problem with both of these reactions is that they use up the free radicals in the system - we'll come back to that problem shortly. The second reaction, of course, also introduces an impurity into the mixture. So what is the useful collision? If a methyl radical hits a bromine molecule (something that's quite likely to occur), the following change can happen: CH3 + Br2 CH3Br + Br

The methyl radical takes one of the bromine atoms to form bromomethane (which is what we want to make), but in the process generates another bromine radical. This new bromine radical can now go through the whole sequence again, and at the end will produce yet another bromine radical - and so on and so on. The process is described as a free radical chain reaction. The chain continues because for every bromine radical that goes in at the beginning, a new one is generated at the end. Chain termination Does this mean that one tiny burst of UV light, splitting one bromine molecule into two free radicals, is enough to convert a whole reactions-worth of methane and bromine into bromomethane and HBr? Sadly, no! As we've seen, there are collisions which result in the removal of free radicals without producing any new ones. These radicals can only be replaced by starting the process all over again with a new burst of light energy. In practice, then, the chains propagate many thousands of times, but eventually any chain will be brought to an end by one of these chain termination processes.

Simplifying all this for exam purposes: The over-all process is known as free radical substitution, or as a free radical chain reaction. Chain initiation The chain is initiated (started) by UV light breaking a bromine molecule into free radicals. Br2 2Br

Chain propagation reactions These are the reactions which keep the chain going. CH4 + Br CH3 + Br2 CH3 + HBr

CH3Br + Br

Chain termination reactions These are reactions which remove free radicals from the system without replacing them by new ones. 2Br CH3 CH3 + Br + CH3 Br2 CH3Br CH3CH3

MULTIPLE SUBSTITUTION IN THE METHANE AND BROMINE REACTION


Warning! We are just about to muddy the water quite considerably! Don't go on until you are sure that you understand the mechanism for the production of bromomethane - and are confident that you could write it in an exam. If you aren't sure about it, go back to that reaction and look at it again. It would be worth checking your syllabus and past exam papers to see if you need to know about these further substitution reactions.

The facts
When a mixture of methane and bromine is exposed to ultraviolet light - typically sunlight - a substitution reaction occurs and the organic product is bromomethane. CH4 + Br2 CH3Br + HBr

However, the reaction doesn't stop there, and all the hydrogens in the methane can in turn be replaced by bromine atoms. That means that you could get any of bromomethane, dibromomethane, tribromomethane or tetrabromomethane. CH4 + Br2 CH3Br + Br2 CH2Br2 + Br2 CHBr3 + Br2 CH3Br + HBr CH2Br2 + HBr CHBr3 + HBr CBr4 + HBr

You might think that you could control which product you got by the proportions of methane and bromine you used, but it isn't as simple as that. If you use enough bromine you will eventually get CBr4, but any other proportions will always lead to a mixture of products.

The mechanisms
The formation of multiple substitution products like di-, tri- and tetrabromomethane can be explained in just the same sort of way as the formation of the original bromomethane. You just have to look at the likely collisions as the reaction progresses. Making dibromomethane You will remember that the over-all equation for the first stage of the reaction is CH4 + Br2 CH3Br + HBr

As the reaction proceeds, the methane is getting used up and bromomethane is taking its place. That means that the argument about what a bromine radical is likely to hit changes during the course of the reaction. As time goes by there is an increasing

chance of it hitting a bromomethane molecule rather than a methane molecule. When that happens, the bromine radical can take a hydrogen from the bromomethane just as well as it could from a methane. In this new case: CH3Br + Br CH2Br + HBr

Notice: The dot representing the electron has been moved against the carbon which is the atom with the unpaired electron. It would be potentially confusing to leave it next to the bromine.

The bromomethyl radical formed can then interact with a bromine molecule in a new propagation step . . . CH2Br + Br2 CH2Br2 + Br

. . . and so dibromomethane is formed and a bromine radical regenerated. These propagation steps continue until the chain is terminated by any two radicals colliding and combining together. Making tri- and tetrabromomethane Obviously, as time goes on, there is an increasing chance of the dibromomethane being hit by a bromine radical - producing these propagation steps giving tribromomethane: CH2Br2 + Br CHBr2 + Br2 CHBr2 + HBr CHBr3 + Br

Care! Don't just skip lightly over these equations. Look carefully at each one so that you understand what is happening, and can relate it to what has gone before. Talk through the equations with yourself. For example: "A bromine radical hits the dibromomethane molecule and steals a hydrogen. That leaves a new radical (I don't know what it's called, but that doesn't really matter, as long as I can work out its formula if I have to!), which then bumps into a bromine

molecule - etc, etc." Doing this helps you to focus properly on the equations. If you just read them quickly, you'll have forgotten all about them again in 15 seconds!

As the amount of tribromomethane builds up, then you will get these steps giving tetrabromomethane: CHBr3 + Br CBr3 + Br2 CBr3 + HBr CBr4 + Br

This is why you will always get a mixture of products whatever the reaction proportions of methane and bromine you use. The whole process is simply governed by chance. Having produced some bromomethane there is no way that you can prevent it from being hit by bromine radicals, and similarly for dibromomethane and tribromomethane. Trying to produce mainly one product If you wanted tetrabromomethane, you could of course get it by using a large excess of bromine, so that eventually all the hydrogens would be replaced. If you wanted mainly bromomethane, you could favour this by using a huge excess of methane so that the chances were always greater of a bromine radical hitting a methane rather than anything else - but even so, you would still get some mixture of products. There is no obvious way of getting mainly dibromomethane or tribromomethane.

SIDE REACTIONS IN THE METHANE AND BROMINE REACTION


You may remember that one of the chain termination steps produces ethane, CH3CH3.

CH3

+ CH3

CH3CH3

If bromine radicals hit that, you are going to get bromoethane and dibromoethane and so on - and in the course of those reactions you will get ethyl radicals which could themselves become involved in chain termination steps leading to propane (from methyl radical hitting ethyl radical) or butane (from two ethyl radicals combining), which could then start to undergo substitution - and on and on! To be honest, all of these side products are going to be present in very small amounts because the reaction producing ethane won't, by chance, happen very often, but it nicely illustrates a typical organic chemistry problem - when you do a reaction in the lab to produce an organic chemical, a high proportion of your time is spent in purifying the product from all the side reactions that have gone on!

THE REACTION BETWEEN METHYLBENZENE AND CHLORINE


A Free Radical Substitution Reaction
This page gives you the facts and a simple, uncluttered mechanism for the free radical substitution reaction between methylbenzene (previously known as toluene) and chlorine. If you want the mechanism explained to you in detail, there is a link at the bottom of the page. Methylbenzene has a methyl group attached to a benzene ring. The hexagon with the circle inside is the standard symbol for this ring. There is a carbon atom at each corner of the hexagon, and a hydrogen atom on each carbon apart from the one with the methyl group attached.
Note: There is no need to worry about the bonding in the benzene ring at this point. If you are interested, you can follow the link - but it isn't important for now.

The facts The reaction we are going to explore happens between methylbenzene and chlorine in the presence of ultraviolet light - typically sunlight. This is a good example of a photochemical reaction - a reaction brought about by light.

Note: These reactions are sometimes described as examples of photocatalysis reactions catalysed by light. It is better to use the term "photochemical" and keep the keep the word "catalysis" for reactions speeded up by actual substances rather than light.

The organic product is (chloromethyl)benzene. The brackets in the name emphasise that the chlorine is part of the attached methyl group, and isn't on the ring. One of the hydrogen atoms in the methyl group has been replaced by a chlorine atom, so this is a substitution reaction. However, the reaction doesn't stop there, and all three hydrogens in the methyl group can in turn be replaced by chlorine atoms. Multiple substitution is dealt with on a separate page, and you will find a link to that at the bottom of this page.
Important! There is another reaction which happens between methylbenzene and chlorine in the absence of light and in the presence of a number of possible catalysts. In that one, substitution happens in the benzene ring instead of in the methyl group. You will find this reaction discussed under electrophilic substitution reactions.

The mechanism The mechanism involves a chain reaction. During a chain reaction, for every reactive species you start off with, a new one is generated at the end - and this keeps the process going.
Species: a useful word which is used in chemistry to mean any sort of particle you want it to mean. It covers molecules, ions, atoms, or (in this case) free radicals.

The over-all process is known as free radical substitution, or as a free radical chain reaction.
Note: If you aren't sure about the words free radical or substitution, read the page What is free radical substitution? Use the BACK button on your browser to return quickly to this page.

Chain initiation

The chain is initiated (started) by UV light breaking a chlorine molecule into free radicals. Cl2 2Cl

Chain propagation reactions These are the reactions which keep the chain going.

Chain termination reactions These are reactions which remove free radicals from the system without replacing them by new ones. If any two free radicals collide, they will join together without producing any new radicals. The simplest example of this is a collision between two chlorine radicals. 2Cl Cl2

EXPLAINING THE REACTION BETWEEN METHYLBENZENE AND CHLORINE

A Free Radical Substitution Reaction This page guides you through the mechanism for the substitution of one of the hydrogen atoms in methylbenzene by one chlorine atom. Multiple substitution is covered separately, and you will find a link at the bottom of the page. We are going to talk through this mechanism in a very detailed way so that you get a feel for what is going on. You couldn't possibly do the same thing in an exam. At the bottom of the page, you will find the condensed down version corresponding to the sort of answer you would produce in an exam. The role of the UV light The ultraviolet light is simply a source of energy, and is being used to break bonds. In fact, the energies in UV are exactly right to break the bonds in chlorine

molecules to produce chlorine atoms.

Note: Only the outer electrons of the chlorine are shown. Notice also that it is quite acceptable to use a simple view of atomic structure. There is no point in using a complicated model of the atom if a simple one will do the job.

Because we want to stress the fact that the chlorine atoms have single unpaired electrons, then we call them chlorine free radicals - or more usually just chlorine radicals. To show that a species (either an atom or a group of atoms) is a free radical, the symbol is written with a dot attached to show the unpaired electron. The splitting of the chlorine molecule would be shown as: Cl2 2Cl

Free radicals are formed if a bond splits evenly - each atom getting one of the two electrons. The name given to this is homolytic fission. What happens to the chlorine radicals? Reactions happen because things hit each other. In this case, you need to think about what the chlorine radicals are likely to hit, and what could happen as a result of that collision. At the moment the mixture contains

lots of methylbenzene molecules lots of chlorine molecules (only a few will have been fractured by the UV light) a few chlorine radicals

Let's start with the unproductive collisions. The least likely collision is between two chlorine radicals. There aren't very many of them in the mixture and so the chances of them hitting each other are relatively small. If they do collide, they will combine to form a chlorine molecule. That's worse than useless because it removes the active free radicals from the system.

2Cl

Cl2

A chlorine radical could also hit a chlorine molecule. If this happens there could possibly be an exchange of chlorine atoms, but nothing new would be formed. It is just a wasted collision. Cl + Cl-Cl Cl-Cl + Cl

Note: There is no difference between the chlorine atoms shown in bold type or ordinary type. They are shown differently so that the exchange is made clear.

The productive collision happens if a chlorine radical hits a methylbenzene molecule.

The chlorine radical removes a hydrogen atom from the methyl group. That hydrogen atom only needs to bring one electron with it to form a new bond to the chlorine, and so one electron is left behind on the carbon atom. A new free radical is formed - called a phenylmethyl radical.
Note: Don't worry about the name of this new radical. All that matters is that you can draw its structure.

What happens to the phenylmethyl radicals? It depends what they collide with. There are three interesting collisions which need to be explored. Two of these involve a set-back to the reaction, and only one is useful. Luckily, the two unhelpful collisions don't happen very often, because they involve collisions between two free radicals - and there won't be many of these present in the mixture at any one time.

A phenylmethyl radical hits a chlorine radical. These will combine to make what you want - (chloromethyl)benzene - but the reaction removes the active free radicals from the system. That stops any further reactions happening.

Even worse, two phenylmethyl radicals could hit each other. Not only does this remove radicals from the system, but produces an unwanted side reaction.

So what is the useful collision? If a phenylmethyl radical hits a chlorine molecule (something that's quite likely to occur), the following change can happen:

The phenylmethyl radical takes one of the chlorine atoms to form (chloromethyl)benzene (which is what we want to make), but in the process generates another chlorine radical. This new chlorine radical can now go through the whole sequence again, and at the end will produce yet another chlorine radical - and so on and so on. The process is described as a free radical chain reaction. The chain continues because for every chlorine radical that goes in at the beginning, a new one is generated at the end. Chain termination Does this mean that one tiny burst of UV light, splitting one chlorine molecule into two free radicals, is enough to convert a whole reactions-worth of methylbenzene and chlorine into (chloromethyl)benzene and HCl? Sadly, no! As we've seen, there are collisions which result in the removal of free radicals without producing any new ones. These radicals can only be replaced by starting the process all over again with a new burst of light energy. In practice, then, the chains propagate many thousands of times, but eventually any chain will be brought to an end by one of these chain termination processes.

Simplifying all this for exam purposes: The over-all process is known as free radical substitution, or as a free radical chain

reaction. Chain initiation The chain is initiated (started) by UV light breaking a chlorine molecule into free radicals. Cl2 2Cl

Chain propagation reactions These are the reactions which keep the chain going.

Chain termination reactions These are reactions which remove free radicals from the system without replacing them by new ones. If any two free radicals collide, they will join together without producing any new radicals. 2Cl Cl2

Important! If you have found this mechanism difficult because of the names and structures involved it would be worth looking at the methane and chlorine reaction. The two mechanisms are identical as far as the substitution is concerned, but the methane / chlorine one looks easier!

MULTIPLE SUBSTITUTION IN THE METHYLBENZENE AND CHLORINE REACTION


Warning! Don't go on until you are sure that you understand the mechanism for the production of (chloromethyl)benzene - and are confident that you could write it in an exam. If you aren't sure about it, go back to that reaction and look at it again. It would be worth checking your syllabus and past exam papers to see if you need to know about these further substitution reactions.

The facts
When a mixture of methylbenzene and chlorine is exposed to ultraviolet light - typically sunlight - a substitution reaction occurs in the methyl group and the organic product is (chloromethyl)benzene.

However, the reaction doesn't stop there, and all the hydrogens in the methyl group can in turn be replaced by chlorine atoms. That means that you could get any of (chloromethyl)benzene, (dichloromethyl)benzene, or (trichloromethyl)benzene.

Care! Look at these equations carefully so that you are sure that you understand what's going on. All that's happening is that the three hydrogens in the methyl group are being replaced by chlorine atoms one at a time.

If you use enough chlorine you will eventually get (trichloromethyl)benzene, but any other proportions will always lead to a mixture of products.

The mechanisms
Making (dichloromethyl)benzene You will remember that the over-all equation for the first stage of the reaction is

As the reaction proceeds, the methylbenzene is getting used up and (chloromethyl)benzene is taking its place. Remember that these reactions happen because chlorine radicals bump into things. As time goes by there is an increasing chance of a chlorine radical hitting a (chloromethyl)benzene molecule rather than a methylbenzene molecule. When that happens, the chlorine radical can take a hydrogen from the (chloromethyl)benzene just as well as it could from a

methylbenzene molecule. In this new case:

The new radical formed can then interact with a chlorine molecule in a new propagation step . . .

. . . and so (dichloromethyl)benzene is formed and a chlorine radical regenerated. These propagation steps continue until the chain is terminated by any two radicals colliding and combining together. Making (trichloromethyl)benzene Obviously, as time goes on, there is an increasing chance of the (dichloromethyl)benzene being hit by a chlorine radical producing these propagation steps giving (trichloromethyl)benzene:

Care! Nothing new is hapening here, but don't just glance briefly at these equations and then move on. Talk them through with yourself. "A chlorine radical takes a hydrogen away from the first molecule (I can't remember what it's called, but that doesn't matter much, because I know how to draw it!) and forms a new radical. That bumps into a chlorine molecule, and gives the product and a new chlorine radical - which can go

through the process all over again. So it's a chain reaction."

You will always get a mixture of products whatever the reaction proportions of methylbenzene and chlorine you use. The whole process is simply governed by chance. Having produced some (chloromethyl)benzene there is no way that you can prevent it from being hit by chlorine radicals, and similarly for (dichloromethyl)benzene.

THE POLYMERISATION OF ETHENE


A Free Radical Addition Reaction
This page gives you the facts and a simple, uncluttered mechanism for the polymerisation of ethene by a free radical addition reaction. If you want the mechanism explained to you in detail, there is a link at the bottom of the page. The facts An addition reaction is one in which two or more molecules join together to give a single product. During the polymerisation of ethene, thousands of ethene molecules join together to make poly(ethene) - commonly called polythene.

The number of molecules joining up is very variable, but is in the region of 2000 to 20000.

Conditions Temperature: about 200C Pressure: about 2000 atmospheres a small amount of oxygen as Initiator: an impurity
Note: The oxygen is sometimes described as a catalyst for the reaction. That's not strictly true. A catalyst can be recovered unchanged at the end of a reaction, but in this case the oxygen is used up. It gets incorporated into the polymer molecules - as you will see shortly.

The mechanism The over-all process is known as free radical addition. Chain initiation The chain is initiated by free radicals, Ra , produced by reaction between some of the ethene and the oxygen initiator. Chain propagation Each time a free radical hits an ethene molecule a new longer free radical is formed.

etc Chain termination Eventually two free radicals hit each other producing a final molecule. The process stops here because no new free radicals are formed.

Because chain termination is a random process, poly(ethene)

will be made up of chains of all sorts of different lengths.

EXPLAINING THE POLYMERISATION OF ETHENE


A Free Radical Addition Reaction
This page guides you through the mechanism for the polymerisation of ethene by a free radical addition reaction. We are going to talk through this mechanism in a very detailed way so that you get a feel for what is going on. You couldn't possibly do the same thing in an exam. At the bottom of the page, you will find the condensed down version corresponding to the sort of answer you would produce in an exam. You will remember that during the polymerisation of ethene, thousands of ethene molecules join together to make poly(ethene) - commonly called polythene. The reaction is done at high pressures in the presence of a trace of oxygen as an initiator.

The function of the oxygen - chain initiation The oxygen reacts with some of the ethene to give an organic peroxide. Organic peroxides are very reactive molecules containing oxygen-oxygen single bonds which are quite weak and which break easily to give free radicals. You can short-cut the process by adding other organic peroxides directly to the ethene instead of using oxygen if you want to. You don't need to worry about the exact formulae of the free radicals which start the reaction off - they vary depending on their source. For simplicity we give them a general formula: Ra

Chain propagation

In an ethene molecule, CH2=CH2, the two pairs of electrons which make up the double bond aren't the same. One pair is held securely on the line between the two carbon nuclei in a bond called a sigma bond. The other pair is more loosely held in an orbital above and below the plane of the molecule known as a pi bond.
Note: It would be helpful - but not essential - if you read about the structure of ethene before you went on. If the diagram above is unfamiliar to you, then you certainly ought to read this background material. Use the BACK button on your browser to return to this page.

Imagine what happens if a free radical approaches the pi bond in ethene.

Note: Don't worry that we've gone back to a simpler diagram. As long as you realise that the pair of electrons shown between the two carbon atoms is in a pi bond - and therefore vulnerable that's all that really matters for this mechanism.

The sigma bond between the carbon atoms isn't affected by any of this. The free radical, Ra , uses one of the electrons in the pi bond to help to form a new bond between itself and the left hand carbon atom. The other electron returns to the right hand carbon.

You can show this using "curly arrow" notation if you want to:

Note: If you aren't sure about about curly arrow notation you can follow this link. Use the BACK button on your browser to return to this page.

This is energetically worth doing because the new bond between the radical and the carbon is stronger than the pi bond which is broken. You would get more energy out when the new bond is made than was used to break the old one. The more energy that is given out, the more stable the system becomes. What we've now got is a bigger free radical - lengthened by CH2CH2. That can react with another ethene molecule in the same way:

So now the radical is even bigger. That can react with another ethene - and so on and so on. The polymer chain gets longer and longer. Chain termination The chain doesn't, however, grow indefinitely. Sooner or later two free radicals will collide together.

That immediately stops the growth of two chains and produces one of the final molecules in the poly(ethene). It is important to realise that the poly(ethene) is going to be a mixture of molecules of different sizes, made in this sort of random way. Simplifying all this for exam purposes

The over-all process is known as free radical addition. Chain initiation The chain is initiated by free radicals, Ra , produced by reaction between some of the ethene and the oxygen initiator. Chain propagation Each time a free radical hits an ethene molecule a new longer free radical is formed.

etc Chain termination Eventually two free radicals hit each other producing a final molecule. The process stops here because no new free radicals are formed.

Because chain termination is a random process, poly(ethene) will be made up of chains of all sorts of different lengths.

HYDROGEN BROMIDE AND ALKENES: THE PEROXIDE EFFECT


This page gives you the facts and simple uncluttered mechanisms for the free radical addition of hydrogen bromide to alkenes - often known as the "peroxide effect". If you want the mechanisms explained to you in more detail, there is a link at the bottom of the page.

Addition to symmetrical alkenes


A symmetrical alkene is one like ethene where the groups at both ends of the carbon-carbon double bond are the same. The facts The reaction happens at room temperature in the presence of organic peroxides or some oxygen from the air. Alkenes react very slowly with oxygen to produce traces of organic peroxides so the two possible conditions are equivalent to each other. The reaction is a simple addition of the hydrogen bromide. For example, with ethene:

With a symmetrical alkene you get exactly the same product in the absence of the organic peroxides or oxygen - but the mechanism is different.

The mechanism Hydrogen halides (hydrogen chloride, hydrogen bromide and the rest) usually react with alkenes using an electrophilic addition mechanism. However, in the presence of organic peroxides, hydrogen bromide adds by a different mechanism.
Note: If you are interested, you will find the electrophilic addition mechanism for the addition of hydrogen bromide and other hydrogen halides to alkenes if you follow this link. You may need to explore several pages in this section. Use the BACK button (or the HISTORY file or GO menu) on your browser to return to this page.

With the organic peroxides present you get a free radical chain

reaction. Chain initiation The chain is initiated by free radicals produced by an oxygenoxygen bond in the organic peroxide breaking.

These free radicals extract a hydrogen atom from a hydrogen bromide molecule to produce bromine radicals.

Chain propagation A bromine radical joins to the ethene using one of the electrons in the pi bond. That creates a new radical with the single electron on the other carbon atom.

That radical reacts with another HBr molecule to produce bromoethane and another bromine radical to continue the process.

etc Chain termination Eventually two free radicals hit each other and produce a molecule of some sort. The process stops here because no new free radicals are formed.

Addition to unsymmetrical alkenes


An unsymmetrical alkene is one like propene where the groups at either end of the carbon-carbon double bond are different.

The facts The reaction happens under the same conditions as with a symmetrical alkene, but there is a complication because the hydrogen and the bromine can add in two different ways. Which way they add depends on whether there are organic peroxides (or oxygen) present or not.

Normally, when a molecule HX adds to a carbon-carbon double bond, the hydrogen becomes attached to the carbon with the more hydrogens on already. This is known as Markovnikov's Rule. Because the HBr adds on the "wrong way around " in the presence of organic peroxides, this is often known as the peroxide effect or anti-Markovnikov addition. In the absence of peroxides, hydrogen bromide adds to propene via an electrophilic addition mechanism. That gives the product predicted by Markovnikov's Rule.

The free radical mechanism Chain initiation This is exactly the same as in the ethene case above.

Chain propagation When the bromine radical joins to the propene, it attaches so that a secondary radical is formed. This is more stable (and so easier to form) than the primary radical which would be formed if it attached to the other carbon atom.

That radical reacts with another HBr molecule to produce 1bromopropane and another bromine radical to continue the process.

etc Chain termination Eventually two free radicals hit each other and produce a molecule of some sort. The process stops here because no new free radicals are formed.

Why don't the other hydrogen halides behave in the same way? The reason that hydrogen bromide adds in an anti-Markovnikov fashion in the presence of organic peroxides is simply a question of reaction rates. The free radical mechanism is much faster than the alternative electrophilic addition mechanism. Both mechanisms happen, but most of the product is the one from the free radical mechanism because that is working faster. With the other hydrogen halides, the opposite is true. Hydrogen fluoride The hydrogen-fluorine bond is so strong that fluorine radicals

aren't formed in the initiation step. Hydrogen chloride With hydrogen chloride, the second half of the propagation stage is very slow. If you do a bond enthalpy sum, you will find that the following reaction is endothermic.

This is due to the relatively high hydrogen-chlorine bond strength. Hydrogen iodide In this case, the first step of the propagation stage turns out to be endothermic and this slows the reaction down. Not enough energy is released when the weak carbon-iodine bond is formed.

In the case of hydrogen bromide, both steps of the propagation stage are exothermic.

EXPLAINING THE "PEROXIDE EFFECT" IN THE REACTION BETWEEN HYDROGEN BROMIDE AND ALKENES
This page guides you through the mechanism for the free radical addition of hydrogen bromide to alkenes - often known as the "peroxide effect".
Note: If you just want the facts and mechanism with a minimum of discussion you will find them by following this link.

Free radical addition to a carbon-carbon double bond


If you have read the introductory page (see above), you will know that hydrogen bromide adds to the carbon-carbon double bond in alkenes via a free radical mechanism in the presence of organic peroxides or oxygen from the air. Oxygen reacts slowly with alkenes to produce small amounts of organic peroxides, so we don't need to look at that as a separate case. We'll start by looking at the general case without worrying about what is attached at either end of the carbon-carbon double bond.

The function of the organic peroxides - chain initiation Organic peroxides are compounds containing an oxygen-oxygen single bond, and are commonly given a general formula R-O-OR. The "R" groups can be quite complicated and aren't necessarily just simple alkyl groups. The oxygen-oxygen bond is quite weak, and breaks easily so that each oxygen gets a single electron. Free radicals are formed.

If these free radicals collide with a hydrogen bromide molecule, a hydrogen atom is transferred, breaking the hydrogen-bromine bond to produce bromine radicals.

Chain propagation In any alkene (like ethene, for example), the two pairs of

electrons which make up the double bond aren't the same. One pair is held securely on the line between the two carbon nuclei in a bond called a sigma bond. The other pair is more loosely held in an orbital above and below the plane of the molecule known as a pi bond.

Note: It would be helpful - but not essential - if you read about the structure of ethene before you went on. If the diagram above is unfamiliar to you, then you certainly ought to read this background material. Use the BACK button on your browser to return to this page.

Imagine what happens if a free radical approaches the pi bond in an alkene. Once again, we'll draw it as if it is ethene - but it would apply to any case. The bromine radical uses one of the electrons in the pi bond to help to form a new bond between itself and the left hand carbon atom. The other electron returns to the right hand carbon.

Note: Don't worry that we've gone back to a simpler diagram. It is perfectly adequate for this discussion.

The sigma bond between the carbon atoms isn't affected by any of this. Now this new free radical reacts with a hydrogen bromide molecule. It takes a hydrogen atom from it, leaving another bromine radical.

The bromine radical can then react with another carbon-carbon double bond, which eventually produces a new bromine radical and so on, and so on . . . There is a chain reaction.

Chain termination The chain will be broken when any two radicals happen to hit each other and form a new bond using both of the single electrons. Removing a free radical from the system without producing a new one immediately stops that particular chain.

What happens if the alkene is unsymmetrical? Unsymmetrical alkenes? An unsymmetrical alkene is one like propene, CH3CH=CH2. At one end of the double bond there is a CH3 group and a hydrogen atom. At the other end there are two hydrogen atoms. A question of orientation The problem with these unsymmetrical alkenes is that you could get two different products depending on which end of the bond the hydrogen and the bromine add. In fact, under free radical conditions, most of the product is 1-

bromopropane.

This happens because when the bromine radical attacks the pi bond, it joins to the carbon atom at the CH2 end of the double bond rather than the CH end.

Why does the bromine add this way? You might think that there would be an equal chance of it attaching to either end, but where it attaches is controlled by the stability of the free radical formed. The more stable radical will be formed more quickly. Think of this in terms of activation energy.

The activation energy will be lower for the reaction where the bromine attaches to the end carbon because the radical produced is more energetically stable. The stability of various sorts of radicals What matters is the number of carbon atoms attached to the carbon with the single electron. Looking at the simplest possibilities:

Tertiary radicals are more stable than secondary ones, and secondary radicals are more stable than primary. In the case of a bromine radical attacking the double bond in propene, it forms a secondary radical rather than a primary one because it is more stable.

Note: The order of stability of the various tertiary, secondary and primary free radicals exactly reflects the order of stability of carbocations (carbonium ions). If you are interested in following this link, use the BACK button on your browser to return to this page.

The rest of the reaction Once the bromine has attached to the carbon to form the secondary radical, there is nothing different about the rest of the reaction. The new radical takes a hydrogen from a hydrogen bromide molecule. This produces the 1-bromopropane and a bromine radical.

The bromine radical now goes into another cycle exactly as before to continue the chain reaction.

ELECTROPHILIC ADDITION MECHANISMS MENU


Addition to symmetrical alkenes
Covers addition to symmetrical alkenes like ethene and cyclohexene. A symmetrical alkene has the same groups attached to both ends of the carbon-carbon double bond. What is electrophilic addition? . . . An explanation of the terms addition and electrophile, together with a general mechanism for these reactions. The reaction with hydrogen halides . . . The mechanism for the reaction between ethene (and cyclohexene) and hydrogen halides (like hydrogen bromide). The reaction with sulphuric acid . . . The mechanism for the reaction between ethene (and cyclohexene) and sulphuric acid. The reaction with bromine . . . The mechanism for the reaction between ethene (and cyclohexene) and bromine.

Addition to unsymmetrical alkenes


Covers addition to unsymmetrical alkenes like propene. An unsymmetrical alkene has different groups attached to each end of the carbon-carbon double bond.
Warning! Don't even think about reading articles in this section until you are sure you understand the corresponding reaction(s) above!

Carbocations (carbonium ions) and their stability . . .

Essential pre-reading before you tackle anything else in this section. Why unsymmetric alkenes are a problem . . . Explains the reasons behind Markovnikov's Rule, and gives a general mechanism for these more awkward reactions. This is also essential reading before you look at specific reactions. The reaction with hydrogen halides . . . The mechanism for the reaction between propene and hydrogen halides (like hydrogen bromide). The reaction with sulphuric acid . . . The mechanism for the reaction between propene and sulphuric acid. The reaction with bromine . . . The mechanism for the reaction between propene and bromine.

ELECTROPHILIC ADDITION
Background
Electrophilic addition happens in many of the reactions of compounds containing carbon-carbon double bonds - the alkenes. The structure of ethene We are going to start by looking at ethene, because it is the simplest molecule containing a carbon-carbon double bond. What is true of C=C in ethene will be equally true of C=C in more complicated alkenes. Ethene, C2H4, is often modelled as shown on the right. The double bond between the carbon atoms is, of course, two pairs of shared electrons. What the diagram doesn't show is that the two pairs aren't the same as each other. One of the pairs of electrons is held on the line between the two carbon

nuclei as you would expect, but the other is held in a molecular orbital above and below the plane of the molecule. A molecular orbital is a region of space within the molecule where there is a high probability of finding a particular pair of electrons. In this diagram, the line between the two carbon atoms represents a normal bond - the pair of shared electrons lies in a molecular orbital on the line between the two nuclei where you would expect them to be. This sort of bond is called a sigma bond. The other pair of electrons is found somewhere in the shaded part above and below the plane of the molecule. This bond is called a pi bond. The electrons in the pi bond are free to move around anywhere in this shaded region and can move freely from one half to the other.
Note: This diagram shows a side view of an ethene molecule. The dotted lines to two of the hydrogens show bonds going back into the screen or paper away from you. The wedge shapes show bonds coming out towards you.

The pi electrons are not as fully under the control of the carbon nuclei as the electrons in the sigma bond and, because they lie exposed above and below the rest of the molecule, they are relatively open to attack by other things.
Note: Check your syllabus to see if you need to know how a pi bond is formed. Haven't got a syllabus? If you are working towards a UK-based exam, find out how to get one by following this link. If you do need to know about the bonding in ethene in detail, follow this link as well.

Electrophiles An electrophile is something which is attracted to electron-rich regions in other molecules or ions. Because it is attracted to a negative region, an electrophile must be something which carries either a full positive charge, or

has a slight positive charge on it somewhere.


Note: The ending ". . phile" means a liking for. For example, a francophile is someone who likes the French; an anglophile is someone who likes the English.

Ethene and the other alkenes are attacked by electrophiles. The electrophile is normally the slightly positive ( +) end of a molecule like hydrogen bromide, HBr.
Note: If you aren't sure about why some bonds are polar, read the page on electronegativity. Use the BACK button on your browser to return to this page.

Electrophiles are strongly attracted to the exposed electrons in the pi bond and reactions happen because of that initial attraction - as you will see shortly. You might wonder why fully positive ions like sodium, Na+, don't react with ethene. Although these ions may well be attracted to the pi bond, there is no possibility of the process going any further to form bonds between sodium and carbon, because sodium forms ionic bonds, whereas carbon normally forms covalent ones. Addition reactions In a sense, the pi bond is an unnecessary bond. The structure would hold together perfectly well with a single bond rather than a double bond. The pi bond often breaks and the electrons in it are used to join other atoms (or groups of atoms) onto the ethene molecule. In other words, ethene undergoes addition reactions. For example, using a general molecule X-Y . . .

Summary: electrophilic addition reactions An addition reaction is a reaction in which two molecules join together to make a bigger one. Nothing is lost in the process. All the atoms in the original molecules are found in the bigger one. An electrophilic addition reaction is an addition reaction which happens because what we think of as the "important" molecule is attacked by an electrophile. The "important" molecule has a region of high electron density which is attacked by something carrying some degree of positive charge.
Note: When we talk about reactions of alkenes like ethene, we think of the ethene as being attacked by other molecules such as hydrogen bromide. Because ethene is the molecule we are focusing on, we quite arbitrarily think of it as the central molecule and hydrogen bromide as its attacker. There's no real justification for this, of course, apart from the fact that it helps to put things in some sort of logical pattern. In reality, the molecules just collide and may react if they have enough energy and if they are lined up correctly.

Understanding the electrophilic addition mechanism


The mechanism for the reaction between ethene and a molecule X-Y It is very unlikely that any two different atoms joined together will have the same electronegativity. We are going to assume that Y is more electronegative than X, so that the pair of electrons is pulled slightly towards the Y end of the bond. That means that the X atom carries a slight positive charge.
Note: Once again, if you aren't sure about electronegativity and bond polarity follow this link

before you read on. Use the BACK button on your browser to return to this page.

The slightly positive X atom is an electrophile and is attracted to the exposed pi bond in the ethene. Now imagine what happens as they approach each other.

You are now much more likely to find the electrons in the half of the pi bond nearest the XY. As the process continues, the two electrons in the pi bond move even further towards the X until a covalent bond is made. The electrons in the X-Y bond are pushed entirely onto the Y to give a negative Y- ion.

Help! Why does the carbon atom have a positive charge? The pi bond was originally made using

an electron from each carbon atom, but both of these electrons have now been used to make a bond to the X atom. This leaves the right-hand carbon atom an electron short - hence positively charged. Note also that we are only showing one of the pairs of electrons around the Y ion. There will be other lone pairs as well, but we are only actually interested in the one we've drawn.

Important term An ion in which the positive charge is carried on a carbon atom is called a carbocation or a carbonium ion (an older term).

In the final stage of the reaction the electrons in the lone pair on the Y- ion are strongly attracted towards the positive carbon atom. They move towards it and form a co-ordinate (dative covalent) bond between the Y and the carbon.
Help! A co-ordinate (dative covalent) bond is simply a covalent bond in which both shared electrons originate from the same atom. The bond formed between the X and the other carbon atom was also a co-ordinate bond. Once a coordinate bond has been formed there is no difference whatsoever between it and any other covalent bond.

How to write this mechanism in an exam The movements of the various electron pairs are shown using curly arrows.
Help! If you aren't sure about the use of curly arrows in mechanisms, you must follow this link before you go on.

Use the BACK button on your browser to return to this page.

Don't leave this page until you are sure that you understand how this relates to the electron pair movements drawn in the previous diagrams.

THE REACTION BETWEEN SYMMETRICAL ALKENES AND THE HYDROGEN HALIDES


This page gives you the facts and a simple, uncluttered mechanism for the electrophilic addition reactions between the hydrogen halides and alkenes like ethene and cyclohexene. Hydrogen halides include hydrogen chloride and hydrogen bromide. If you want the mechanisms explained to you in detail, there is a link at the bottom of the page.

Electrophilic addition reactions involving hydrogen bromide


The facts Alkenes react with hydrogen bromide in the cold. The double bond breaks and a hydrogen atom ends up attached to one of the carbons and a bromine atom to the other. In the case of ethene, bromoethane is formed.
Note: Be careful when you write the names of the addition products that you change the ene ending in the original alkene (showing the C=C) into an ane ending

(showing that it has been replaced by C-C).

With cyclohexene you get bromocyclohexane.

The structures of the cyclohexene and the bromocyclohexane are often simplified:

Note: Each corner in one of these diagrams represents a carbon atom. Each carbon atom has enough hydrogens attached to make the total number of bonds up to 4. In the case of the bromocyclohexane, it isn't necessary to write the new hydrogen into the diagram, but it is helpful to put it there to emphasise that addition has happened.

Be sure that you understand the relationship between these simplified diagrams and the full structures. The mechanisms

The reactions are examples of electrophilic addition. With ethene and HBr:

and with cyclohexene:

Electrophilic addition reactions involving the other hydrogen halides


The facts Hydrogen chloride and the other hydrogen halides add on in exactly the same way. For example, hydrogen chloride adds to ethene to make chloroethane:

The only difference is in how fast the reactions happen with the different hydrogen halides. The rate of reaction increases as you go from HF to HCl to HBr to HI. HF HCl HBr HI fastest reaction slowest reaction

The reason for this is that as the halogen atoms get bigger, the strength of the hydrogen-halogen bond falls. Bond strengths (measured in kilojoules per mole) are:

H-F H-Cl H-Br H-I

568 432 366 298


Note: You may find slightly different values depending on which data source you use. It doesn't matter - the differences are minor and the pattern is always the same.

As you have seen in the HBr case, in the first step of the mechanism the hydrogen-halogen bond gets broken. If the bond is weaker, it will break more readily and so the reaction is more likely to happen. The mechanisms The reactions are still examples of electrophilic addition. With ethene and HCl, for example:

This is exactly the same as the mechanism for the reaction between ethene and HBr, except that we've replaced Br by Cl. All the other mechanisms for symmetrical alkenes and the hydrogen halides would be done in the same way.

EXPLAINING THE REACTION BETWEEN SYMMETRICAL ALKENES AND THE HYDROGEN HALIDES

This page guides you through the mechanism for the electrophilic addition of hydrogen halides such as hydrogen bromide with symmetrical alkenes like ethene or cyclohexene. Unsymmetrical alkenes are covered separately, and you will find a link at the bottom of the page.

Electrophilic addition reactions involving hydrogen bromide


Hydrogen bromide is chosen as a typical hydrogen halide. Bromine is more electronegative than hydrogen. That means that the bonding pair of electrons is pulled towards the bromine end of the bond, and so the hydrogen bromide molecule is polar.
Note: If you aren't sure about electronegativit y and bond polarity follow this link before you read on. Use the BACK button on your browser to return to this page.

The reaction of ethene with hydrogen bromide The structure of ethene is shown in the diagram on the right. The pi bond is an orbital above and below the plane of the rest of the molecule, and relatively exposed to things around it. The two electrons in this orbital are highly attractive to anything which is positively charged.

Note: If you aren't sure about this, it would be a

good idea to read the introductory page on electrophilic addition before you go on. Use the BACK button on your browser to return to this page.

The slightly positive hydrogen atom in the hydrogen bromide acts as an electrophile, and is strongly attracted to the electrons in the pi bond.
Electrophile: A substance with a strong attraction to a negative region in another substance. Electrophiles are either fully positive ions, or the slightly positive end of a polar molecule.

The electrons from the pi bond move down towards the slightly positive hydrogen atom. In the process, the electrons in the H-Br bond are repelled down until they are entirely on the bromine atom, producing a bromide ion.
Help! If you aren't sure about the use of curly arrows in

mechanisms, you must follow this link before you go on. Use the BACK button on your browser to return to this page.

That leaves you with these two ions at this half-way stage of the reaction:

The ion with a positive charge on the carbon atom is called a carbocation or carbonium ion (an older term). Why is there a positive charge on the carbon atom? The pi bond was originally made up of an electron from each of the carbon atoms. Both of those electrons have been used to make a new bond to the hydrogen. That leaves the right-hand carbon an electron short - hence positively charged. In the second stage of the mechanism, the lone pair of electrons on the bromide ion is strongly attracted to the positive carbon and moves towards it until a bond is formed.

Note: For clarity only one of the lone pairs around the bromide ion is shown. That's perfectly acceptable,

because the other three lone pairs aren't involved in the process they are pointing in the wrong directions.

The overall mechanism is therefore

The reaction of cyclohexene with hydrogen bromide Cyclohexene is chosen an an example of a fairly commonly used symmetrical alkene. The fact that it is a ring structure doesn't make any difference to the mechanism. The full structure of cyclohexene is

but it is often abbreviated to

In this diagram, there is a carbon atom at each corner, and enough hydrogens attached to each carbon to bring the total number of bonds per carbon atom up to 4. The double bond has an easily attacked pi bond exactly as in ethene, and the

electrons in that bond are attracted towards the slightly positive hydrogen atom in the HBr. Once again, the pi bond electrons swing to make a bond with the hydrogen, and push the electrons in the H-Br bond fully onto the bromine, making a bromide ion.

Care! Think carefully about which way the pi bond electrons swing. In this case, think of them as being pivotted about the top carbon atom. It is therefore that carbon atom which is joined to the new hydrogen.

In the second stage, one of the lone pairs of electrons on the bromide ion is attracted to the positively charged carbon atom and forms a bond with it.

The overall mechanism is therefore:

Electrophilic addition reactions involving the other hydrogen halides

The mechanisms The other hydrogen halides behave in exactly the same way as hydrogen bromide. For example, compare the reaction between ethene and hydrogen bromide with the one between ethene and hydrogen chloride.

There's no need to learn both mechanisms. As long as you know one of them, all you have to do is swap one halogen atom for another. That's equally true for hydrogen fluoride or hydrogen iodide. The different rates of reaction The rate of reaction increases as you go from HF to HCl to HBr to HI. HF HCl HBr HI fastest reaction slowest reaction

The reason for this is that as the halogen atoms get bigger, the strength of the hydrogen-halogen bond falls. Bond strengths (measured in kilojoules per mole) are: H-F H-Cl H-Br 568 432 366

H-I

298

In the first step of these mechanisms, the hydrogen-halogen bond breaks as the electron pair is forced down onto the halogen atom. Breaking bonds needs energy, and if the bond is weaker, it will break more easily - needing less energy. That means that the activation energy for the reactions will fall as you go from hydrogen fluoride to hydrogen iodide. The lower the activation energy, the faster the reaction.
Activation energy: The minimum energy needed before a reaction will occur. In this case it is the energy needed to break the various bonds and make the carbocation and the halide ion.

Beware! A red herring! People sometimes get confused because there is another tempting place to look for the reason why the reaction rates are different between the various hydrogen halides. The halogens have different electronegativities - with fluorine being the most electronegative and iodine the least. That means that the hydrogen in HF will have the greatest positive charge and so will be attracted most strongly to the pi bond. It would be tempting to think that that would produce the fastest reaction - but not so! Although the HF may well be attracted most strongly, attraction alone isn't enough. If anything is to happen, bonds have to be broken - and here HF is at a disadvantage, because the bond is very strong.

The lesson from all this When you are trying to find reasons for differing rates of reactions, always look first at differences in bond strengths. Electronegativity differences may be interesting, but rarely give you the answer you want!

THE REACTION BETWEEN UNSYMMETRICAL ALKENES AND THE HYDROGEN HALIDES


This page gives you the facts and a simple, uncluttered mechanism for the electrophilic addition reactions between the hydrogen halides and alkenes like propene. Hydrogen halides include hydrogen chloride and hydrogen bromide. If you want the mechanisms explained to you in detail, there is a link at the bottom of the page. An unsymmetrical alkene is one like propene in which the groups or atoms attached to either end of the carbon-carbon double bond are different. For example, in propene there are a hydrogen and a methyl group at one end, but two hydrogen atoms at the other end of the double bond. But-1-ene is another unsymmetrical alkene.

Electrophilic addition reactions involving hydrogen bromide


The facts As with all alkenes, unsymmetrical alkenes like propene react with

hydrogen bromide in the cold. The double bond breaks and a hydrogen atom ends up attached to one of the carbons and a bromine atom to the other. In the case of propene, 2-bromopropane is formed.

This would normally be written in a more condensed form as

The product is 2-bromopropane.


Note: There is another possible reaction between unsymmetrical alkenes and hydrogen bromide (but not the other hydrogen halides) unless the hydrogen bromide and alkene are absolutely pure. A different mechanism happens (a free radical chain reaction not on UK A' level syllabuses) which leads to the hydrogen and bromine adding the opposite way round. For A' level purposes, you don't need to worry about that. However, if you are interested, you will find the free radical addition mechanism by following this link. Use the BACK button on your browser to return to this page later.

This is in line with Markovnikov's Rule which says:

When a compound HX is added to an unsymmetrical alkene, the hydrogen becomes attached to the carbon with the most hydrogens attached to it already. In this case, the hydrogen becomes attached to the CH2 group, because the CH2 group has more hydrogens than the CH group. Notice that only the hydrogens directly attached to the carbon atoms at either end of the double bond count. The ones in the CH3 group are totally irrelevant.
Warning! Markovnikov's Rule is a useful guide for you to work out which way round to add something across a double bond, but it isn't the reason why things add that way. As a general principle, don't quote Markovnikov's Rule in an exam unless you are specifically asked for it.

The mechanism This is an example of electrophilic addition.

The addition is this way around because the intermediate carbocation (previously called a carbonium ion) formed is secondary. This is more stable (and so is easier to form) than the primary carbocation which would be produced if the hydrogen became attached to the centre carbon atom and the bromine to the end one.

Electrophilic addition reactions involving the other hydrogen halides

The facts Hydrogen fluoride, hydrogen chloride and hydrogen iodide all add on in exactly the same way as hydrogen bromide. The only differences lie in the rates of reaction: HF HCl HBr HI fastest reaction slowest reaction

This is because the hydrogen-halogen bond gets weaker as the halogen atom gets bigger. If the bond is weaker, it breaks more easily and so the reaction is faster. If the halogen is given the symbol X, the equation for the reaction with propene is:

Notice that the product is still in line with Markovnikov's Rule. The mechanism These are still examples of electrophilic addition. Again using X to stand for any halogen:

Again, the intermediate carbocation formed is secondary. This is more stable than the primary carbocation ion which would be formed if the hydrogen attached to the centre carbon atom and the

X to the end one. If it is more stable it will be easier to make.

EXPLAINING THE REACTION BETWEEN UNSYMMETRICAL ALKENES AND THE HYDROGEN HALIDES
This page guides you through the mechanism for the electrophilic addition of hydrogen halides such as hydrogen bromide to unsymmetrical alkenes like propene.
Important! To make sense of this page, you will need to understand about the structure and stability of carbocations (previously called carbonium ions) and be confident about electrophilic addition to simple alkenes like ethene. If you aren't sure about either of these things, follow these links first. You would also find it easier if you first read about the electrophilic addition

reactions between hydrogen halides and symmetrical alkenes like ethene, and addition to unsymmetrical alkenes in general. If you have just come from those, ignore these links!

Electrophilic addition reactions involving hydrogen bromide


If you want the mechanism for one of the other hydrogen halides, simply replace Br by whatever else you are interested in - F or Cl or I. There is no difference whatsoever in the mechanisms. You might, however, need to be aware that there is an alternative mechanism involving hydrogen bromide and alkenes if the reaction mixture is impure in the presence of organic peroxides or oxygen from the air.
Note: The different mechanism (a free radical chain reaction not on UK A' level syllabuses) leads to the hydrogen and bromine adding the opposite way round. For A' level purposes, you don't need to worry about

that. However, if you are interested, you will find the free radical addition mechanism by following this link. Use the BACK button on your browser to return to this page later.

Hydrogen bromide as an electrophile Hydrogen bromide is chosen as a typical hydrogen halide. Bromine is more electronegative than hydrogen. That means that the bonding pair of electrons is pulled towards the bromine end of the bond, and so the hydrogen bromide molecule is polar.
Note: If you aren't sure about electronegativit y and bond polarity follow this link before you read on. Use the BACK button on your browser to return to this page.

The slightly positive hydrogen atom will be attracted to negative regions in other molecules, and is

therefore an electrophile.
Electrophile: A substance with a strong attraction to a negative region in another substance. Electrophiles are either fully positive ions, or the slightly positive end of a polar molecule.

The reaction of propene with hydrogen bromide The double bond in all alkenes is made up of two different parts. One pair of electrons lies on the line between the two nuclei where you would expect them to be. This is called a sigma bond. The other pair lies in an orbital above and below the plane of the rest of the molecule, and is called a pi bond. The pi bond is weaker than a sigma bond and is very vulnerable to attack.
Note: If this isn't fairly obvious to you, you really ought to read the page introducing electrophilic addition before you go on. Use the BACK button on your browser to return to this page.

As the HBr approaches the pi bond, the electrons in that bond are attracted towards the slightly positive hydrogen atom. That repels the electrons in the hydrogen-bromine bond down towards the bromine.

The electron movements continue until a new bond is made between one of the carbon atoms and the hydrogen. The bromine now has both electrons from the H-Br bond, and so is negatively charged as a bromide ion. The problem is that there are two possible ways that the pi bond electrons could move. They could form a bond between the hydrogen and the left-hand carbon:

or they could form a bond with the right-hand one:

It's the second of these changes that happens more readily. In that case, a secondary carbocation is formed - and that's more energetically stable than the primary one formed in the first possibility. Because the secondary ion is more energetically stable, it will form more easily and so the reaction needs less activation energy.
Important! If you don't understand about the structure and stability of carbocations (carbonium ions) follow this link. Activation energy: The minimum energy needed before a reaction will occur. In this case it is the energy needed to break the various bonds and make the carbocation and the bromide ion.

Once the ions have been formed, the lone pair on the bromide ion is strongly attracted towards the positive

carbon atom. It moves towards it and forms a bond.

Note: There are actually 4 lone pairs around the bromide ion, but we are only interested in the one shown.

That leaves you with the over-all mechanism:

Important! If you've had problems with this page you might find it useful to read about addition to unsymmetrical alkenes in general, and then come back here again.

THE REACTION BETWEEN SYMMETRICAL ALKENES AND SULPHURIC ACID

This page gives you the facts and a simple, uncluttered mechanism for the electrophilic addition reactions between sulphuric acid and alkenes like ethene and cyclohexene. If you want the mechanisms explained to you in detail, there is a link at the bottom of the page.

The electrophilic addition reaction between ethene and sulphuric acid


The facts Alkenes react with concentrated sulphuric acid in the cold to produce alkyl hydrogensulphates. Ethene reacts to give ethyl hydrogensulphate.

The structure of the product molecule is sometimes written as CH3CH2HSO4, but the version in the equation is better because it shows how all the atoms are linked up. You may also find it written as CH3CH2OSO3H. Confused by all this? Don't be! All you need to do is to learn the structure of sulphuric acid, and after that the mechanism is exactly the same as the one with hydrogen bromide. As you will find out, the formula of the product follows from the mechanism in an inevitable way.
Important! Learn this structure for sulphuric acid. Sketch it over and over again until you can't possibly get it wrong.

The mechanism for the reaction between ethene and sulphuric acid

Sulphuric acid as an electrophile The hydrogen atoms are attached to very electronegative oxygen atoms which means that the hydrogens will have a slight positive charge while the oxygens will be slightly negative. In the mechanism, we just focus on one of the hydrogen to oxygen bonds, because the other one is too far from the carbon-carbon double bond to be involved in any way. The mechanism

Look carefully at the structure of the product so that you can see how it relates to the various formulae given earlier (CH3CH2OSO2OH etc).

The electrophilic addition reaction between cyclohexene and sulphuric acid


This time we are going straight for the mechanism without producing an initial equation. This is to show that you can work out the structure of obscure products provided you can write the mechanism. The mechanism for the reaction between cyclohexene and sulphuric acid

Having worked out the structure of the product, you could then write a simple equation for the reaction if you wanted to.

EXPLAINING THE REACTION BETWEEN SYMMETRICAL ALKENES AND SULPHURIC ACID


This page guides you through the mechanism for the electrophilic addition of sulphuric acid to symmetrical alkenes like ethene or cyclohexene. Unsymmetrical alkenes are covered separately, and you will find a link at the bottom of the page.

The electrophilic addition reaction between ethene and sulphuric acid

This reaction looks more complicated than the reaction between ethene and hydrogen bromide, but it isn't! The only problem is that H2SO4 is a more complicated structure than HBr. The mechanisms are exactly the same.
Important! If you aren't sure about the reaction of ethene with HBr follow this link before you read on.

The structure of sulphuric acid Compare the structure of sulphuric acid with that of hydrogen bromide:

We are focussing on only one of the hydrogens in the sulphuric acid because the other one will be pointing away from the double bond in the alkene as the molecules approach each other. In each case, the hydrogen is attached to a more electronegative element, and so carries a slight positive charge. That means that the hydrogen atoms will serve as electrophiles.
Electrophile: A substance with a strong attraction to a negative region in another substance. Electrophiles are either fully positive ions, or the slightly positive end of a polar molecule. If you aren't sure about electronegativity and polar bonds follow this link before

you read on.

When the sulphuric acid reacts, the whole of the shaded part of the molecule remains as a complete unit. What happens to that unit is exactly the same as happens to the bromine in the reaction involving HBr. When you write the mechanisms involving sulphuric acid, keep that shaded part unchanged throughout - apart from where you would change the bromine. For example, you will need to put a lone pair and a negative charge on the oxygen atom in the middle of the mechanism. That's exactly what you had to do with the bromine in the HBr case. The mechanism The structure of ethene is shown in the diagram on the right. The pi bond is an orbital above and below the plane of the rest of the molecule, and relatively exposed to things around it. The two electrons in this orbital are highly attractive to anything which is positively charged.
Note: If you aren't sure about this, it would be a good idea to read the introductory page on electrophilic addition before you go on. Use the BACK button on your browser to return to this page.

The slightly positive hydrogen atom in the sulphuric acid acts as an electrophile, and is strongly attracted to the electrons in the pi bond. The electrons from the pi bond move down towards the slightly

positive hydrogen atom. In the process, the electrons in the hydrogen-oxygen bond are repelled down until they are entirely on the oxygen atom, producing a negative ion. So the first stage of the reaction is:

Help! If you aren't sure about the use of curly arrows in mechanisms, you must follow this link before you go on. Use the BACK button on your browser to return to this page.

The ion with a positive charge on the carbon atom is called a carbocation or carbonium ion (an older term). Why is there a positive charge on the carbon atom? The pi bond was originally made up of an electron from each of the carbon atoms. Both of those electrons have been used to make a new bond to the hydrogen. That leaves the right-hand carbon an electron short - hence positively charged. In the second stage of the mechanism, the lone pair of electrons on the oxygen atom is strongly attracted to the positive carbon and moves towards it until a bond is formed.

Note: There are other lone pairs around the oxygen atom as well, but we are only showing one of them for clarity.

The overall mechanism is therefore

The electrophilic addition reaction between cyclohexene and sulphuric acid


Once again

the pi bond breaks and the pair of electrons is used to form a bond with the hydrogen atom; the electrons in the hydrogen-oxygen bond are pushed on to the oxygen atom giving it a full negative charge; the lower carbon atom in the original C=C bond becomes positively charged because the electron it originally supplied to the pi bond has been moved away to form the new bond.

Note: Be prepared to draw the sulphuric acid various ways around (on its side, upside-down, etc) so that it fits more tidily into the mechanism you are writing. Also: Be careful to attach the hydrogen to the correct carbon atom. As the curly arrow has been drawn, you can think of the electron pair pivotting around the top carbon atom. The electrons stay attached to that carbon, and so that's the one the hydrogen must join on to.

In the second stage, the lone pair on the negatively charged oxygen is attracted towards the positively charge carbon and forms a bond with it.

The overall mechanism is therefore

THE REACTION BETWEEN UNSYMMETRICAL ALKENES AND SULPHURIC ACID


This page gives you the facts and a simple, uncluttered mechanism for the electrophilic addition reactions between sulphuric acid and unsymmetrical alkenes like propene. If you want the mechanism explained to you in detail, there is a link at the bottom of the page.

The electrophilic addition reaction between propene and sulphuric acid


The facts Alkenes react with concentrated sulphuric acid in the cold to produce alkyl hydrogensulphates. In the case of propene, the equation is:

Note: If you are confused about the structure of the product, a similar structure is described in more detail in the page about the reaction between sulphuric acid and symmetrical alkenes.

This is in line with Markovnikov's Rule which says: When a compound HX is added to an unsymmetrical alkene, the hydrogen becomes attached to the carbon with the most hydrogens attached to it already. In this case, the hydrogen becomes attached to the CH2 group, because the CH2 group has more hydrogens than the CH group. Notice that only the hydrogens directly attached to the carbon atoms at either end of the double bond count. The ones in the CH3 group are totally irrelevant. The mechanism This is an example of electrophilic addition.

The addition is this way around because it is easier to produce the secondary carbocation (carbonium ion) than the primary one which would be formed if the hydrogen became attached to the centre carbon atom and the rest of the sulphuric acid to the end one.

EXPLAINING THE REACTION BETWEEN UNSYMMETRICAL ALKENES AND SULPHURIC ACID

This page guides you through the mechanism for the electrophilic addition of sulphuric acid to unsymmetrical alkenes like propene.

The electrophilic addition reaction between propene and sulphuric acid


Important! The reaction between propene and sulphuric acid has exactly the same mechanism as the one involving hydrogen bromide. It just looks more complicated because of the structure of sulphuric acid. Make sure that you understand the mechanism for the reaction between propene and HBr before you read on. It is also essential that you read about the reaction between ethene and sulphuric acid before you worry about the additional problem caused by an unsymmetrical alkene like propene.

The structure of sulphuric acid Compare the structure of sulphuric acid with that of hydrogen bromide:

We are focussing on only one of the hydrogens in the sulphuric acid because the other one will be pointing away from the double bond in the alkene as the molecules approach each other. In each case, the hydrogen is attached to a more electronegative element, and so carries a slight positive charge. That means that the hydrogen atoms will serve as electrophiles. When the sulphuric acid reacts, the whole of the shaded part of the molecule remains as a complete unit. What happens to that unit is

exactly the same as happens to the bromine in the reaction involving HBr. The mechanism Remember that the active part of the double bond is the pi bond which lies in an orbital above and below the plane of the rest of the molecule. The pi bond is very vulnerable to attack.
Note: If this isn't fairly obvious to you, you really ought to read the page introducing electrophilic addition before you go on. Use the BACK button on your browser to return to this page.

The slightly positive hydrogen atom in the sulphuric acid acts as an electrophile, and is strongly attracted to the electrons in the pi bond. As the sulphuric acid approaches the pi bond, the electrons in that bond are drawn down towards the slightly positive hydrogen atom. That repels the electrons in the hydrogen-oxygen bond down towards the oxygen.

The electron movements continue until a new bond is made between one of the carbon atoms and the hydrogen. The oxygen now has both electrons from the H-O bond, and so becomes negatively charged. The problem is that there are two possible ways that the pi bond

electrons could move. They could form a bond between the hydrogen and the left-hand carbon:

or they could form a bond with the right-hand one:

It's the second of these changes that happens more readily. In that case, a secondary carbocation is formed - and that's more energetically stable than the primary one formed in the first possibility. Because the secondary ion is more energetically stable, it will form more easily and so the reaction needs less activation energy - and so happens faster.
Important! If you don't understand about the structure and stability of carbocations (previously known as carbonium ions) follow this link.

Once the ions have been formed, the lone pair on the negative oxygen is strongly attracted towards the positive carbon atom. It moves towards it and forms a bond.

That leaves you with the over-all mechanism:

THE REACTION BETWEEN SYMMETRICAL ALKENES AND BROMINE


This page gives you the facts and a simple, uncluttered mechanism for the electrophilic addition reactions between bromine (and the other halogens) and alkenes like ethene and cyclohexene. If you want the mechanisms explained to you in detail, there is a link at the bottom of the page.

The electrophilic addition of bromine to ethene


The facts Alkenes react in the cold with pure liquid bromine, or with a solution of bromine in an organic solvent like tetrachloromethane. The double bond breaks, and a bromine atom becomes attached to each carbon. The

bromine loses its original red-brown colour to give a colourless liquid. In the case of the reaction with ethene, 1,2-dibromoethane is formed.

This decolourisation of bromine is often used as a test for a carbon-carbon double bond. If an aqueous solution of bromine is used ("bromine water"), you get a mixture of products. The presence of the water complicates the mechanism beyond what is required by current UK A level (or equivalent) syllabuses. The other halogens, apart from fluorine, behave similarly. (Fluorine reacts explosively with all hydrocarbons - including alkenes - to give carbon and hydrogen fluoride.) If you are interested in the reaction with, say, chlorine, all you have to do is to replace Br by Cl in all the equations on this page. The mechanism for the reaction between ethene and bromine The reaction is an example of electrophilic addition.
Warning! There are two versions of the ethene / bromine mechanism in common use, and you must know which your examiners will accept. One version is simplified to bring it into line with the other alkene electrophilic addition mechanisms. You will probably find that your examiners will accept this one, but you must find out to be sure. You almost certainly won't be able to tell this from your syllabus. You need to refer to recent mark schemes, or to any support material that your examiners provide. If you still aren't sure, contact your examiners direct. If you are working towards a UK-based exam, you can find out how to do this by using the link to your Board's web site on the

syllabuses page. The person you need to contact will probably have the title Subject Officer for Chemistry or something similar. Ask whether they want the mechanism for the reaction between bromine and alkenes which proceeds via a carbocation or via a bromonium ion intermediate.

Bromine as an electrophile The bromine is a very "polarisable" molecule and the approaching pi bond in the ethene induces a dipole in the bromine molecule. If you draw this mechanism in an exam, write the words "induced dipole" next to the bromine molecule - to show that you understand what's going on. The simplified version of the mechanism
Note: Use this version unless your examiners insist on the more accurate one.

The more accurate version of the mechanism


Note: Don't learn this unless you have to. There is a real risk of getting confused. If your examiners are happy to accept the simple version, there's no point in making life difficult for yourself.

In the first stage of the reaction, one of the bromine atoms becomes attached to both carbon atoms, with the positive charge being found on the bromine atom.

A bromonium ion is formed.

The bromonium ion is then attacked from the back by a bromide ion formed in a nearby reaction.

The electrophilic addition of bromine to cyclohexene


The facts Cyclohexene reacts with bromine in the same way and under the same conditions as any other alkene. 1,2-dibromocyclohexane is formed.

The mechanism for the reaction between cyclohexene and bromine The reaction is an example of electrophilic addition.
Warning! Again, there are two versions of this mechanism in common use, and you must know which your examiners will accept.

Bromine as an electrophile Again, the bromine is polarised by the approaching pi bond in the cyclohexene. Don't forget to write the words "induced dipole" next to the bromine molecule. The simplified version of the mechanism
Note: Use this version unless your examiners insist on the more accurate one.

The alternative version of the mechanism


Note: Don't learn this unless you have to. If your examiners are happy to accept the simple version, there's no point in making life difficult for yourself.

In the first stage of the reaction, one of the bromine atoms becomes attached to both carbon atoms, with the positive charge being found on the bromine atom. A bromonium ion is formed.

The bromonium ion is then attacked from the back by a bromide ion formed in a nearby reaction.

EXPLAINING THE REACTION BETWEEN SYMMETRICAL ALKENES AND BROMINE


This page guides you through the mechanism for the electrophilic addition of bromine to symmetrical alkenes like ethene or cyclohexene. Unsymmetrical alkenes are covered separately, and you will find a link at the bottom of the page.

The electrophilic addition of bromine to ethene


The structure of ethene The structure of ethene is shown in the diagram on the right. The pi bond is an orbital above and below the plane of the rest of the molecule, and relatively exposed to things around it.

Note: If you aren't sure about this, then you should read the page What is electrophilic addition? before you go on. Use the BACK button on your browser to return to this page.

Bromine as an electrophile Since two identical bromine atoms are joined together in the bromine molecule there is no reason why one atom should pull the bonding pair of electrons towards itself - they must be equally electronegative

and so there won't be any separation of charge, + or -. How, then, can bromine be an electrophile?
Note: If you aren't sure about electronegativity and bond polarity follow this link before you read on. Equally, if you aren't sure about terms like electrophile, then it really would be a good idea to read the page What is electrophilic addition? before you go on. Use the BACK button on your browser to return to this page.

In fact, bromine is a very polarisable molecule - in other words, the electrons in the bond are very easily pushed to one end or the other. As the bromine molecule approaches the ethene, the electrons in the pi bond tend to repel the electrons in the bromine-bromine bond, leaving the nearer bromine slightly positive and the further one slightly negative.

The bromine molecule therefore acquires an induced dipole which is automatically lined up the right way round for a successful attack on the ethene.
Help! What is an "induced dipole"? A dipole is simply a separation of charge between + at one end and - at the other. "Induced" means that it has been created by some external influence (in this case the approach of the pi bond), and didn't already exist. Where it does already exist - as, for example, in HBr - it is called a permanent dipole.

The simplified version of the mechanism


Note: Use this version unless your examiners insist on the more accurate one. If you've come into this web site from a search engine directly to this page, read the notes on the introductory page to this reaction before you go any further.

The electrons from the pi bond move down towards the slightly positive bromine atom.

In the process, the electrons in the Br-Br bond are repelled down until they are entirely on the bottom bromine atom, producing a bromide ion.
Help! If you aren't sure about the use of curly arrows in mechanisms, you must follow this link before you go on. Use the BACK button on your browser to return to this page.

The ion with a positive charge on the carbon atom is called a carbocation or carbonium ion. Why is there a positive charge on the carbon atom? The pi bond was originally made up of an electron from each of the carbon atoms. Both of those electrons have been used to make a new bond to the bromine. That leaves the right-hand carbon an electron short - hence positively charged. In the second stage of the mechanism, the lone pair of electrons on the bromide ion is strongly attracted to the positive carbon and

moves towards it until a bond is formed.

The overall mechanism is therefore

The more accurate version of the mechanism


Note: Don't learn this unless you have to. There is a real risk of getting confused. If your examiners are happy to accept the simple version, there's no point in making life difficult for yourself.

The reaction starts off just the same as in the simplified version, with the pi bond electrons moving down towards the slightly positive bromine atom. But this time, the top bromine atom becomes attached to both carbon atoms, with the positive charge being found on the bromine rather than on one of the carbons. A bromonium ion is formed.

The bromonium ion is then attacked from the back by a bromide ion formed in a nearby reaction. It can't be attacked by its original bromide ion because the bromonium ion is completely cluttered up with a positive bromine on that side.

It doesn't matter which of the carbon atoms the bromide ion attacks the end result would be just the same.
Note: You can't really draw this mechanism tidily in one line because the bromide ion has to be in a different place at the beginning of the second stage than it was at the end of the first stage.

The electrophilic addition of bromine to cyclohexene


The simplified version of the mechanism
Note: Use this version unless your examiners insist on the more accurate one.

The electrons from the pi bond move towards the slightly positive bromine atom.

In the process, the electrons in the bromine-bromine bond are repelled until they are entirely on the right-hand bromine atom, producing a bromide ion. Exactly as with ethene, a carbocation is formed. The bottom carbon atom lost one of its electrons when the pi bond swung towards the bromine. In the second stage of the mechanism, the lone pair of electrons on

the bromide ion is strongly attracted to the positive carbon and moves towards it until a bond is formed.

The overall mechanism is therefore

The alternative version of the mechanism


Note: Don't learn this unless your examiners insist on it. Keep life simple!

The reaction starts off just the same as in the simplified version, with the pi bond electrons moving towards the slightly positive bromine atom. But this time, the left-hand bromine atom becomes attached to both carbon atoms, with the positive charge being found on the bromine rather than on one of the carbons. A bromonium ion is formed.

The bromonium ion is then attacked from the back by a bromide ion formed in a nearby reaction. It can't be attacked by its original bromide ion because approach from that side is hindered by the positive bromine atom.

It doesn't matter which of the carbon atoms on either end of the original double bond the bromide ion attacks - the end result would be just the same.
Note: Once again, you can't really draw this mechanism tidily in one line because of the need to move the bromide ion.

THE REACTION BETWEEN UNSYMMETRICAL ALKENES AND BROMINE


Important! This page assumes that you have already read the page on the addition of bromine to symmetrical alkenes. If you haven't, you must read it before you go on. It contains important advice that you will need to make best use of this page.

This page gives you the facts and a simple, uncluttered mechanism for the electrophilic addition reactions between bromine and alkenes like propene. If you want the mechanism explained to you in detail, there is a link at the bottom of the page. An unsymmetrical alkene is one like propene in which the groups or atoms attached to either end of the carbon-carbon double bond are different. For example, in propene there are a hydrogen and a methyl group at one end, but two hydrogen atoms at the other end of the double bond. But-1-ene is another unsymmetrical alkene.

The electrophilic addition of bromine to propene


The facts In common with all other alkenes, propene reacts in the cold with pure liquid bromine, or with a solution of bromine in an organic solvent like tetrachloromethane. The double bond breaks, and a bromine atom becomes attached to each carbon. The bromine loses its original red-brown colour to give a colourless liquid. In the case of the reaction with propene, 1,2-dibromopropane is formed.

The other halogens, apart from fluorine, behave similarly. (Fluorine reacts explosively with all hydrocarbons - including alkenes - to give carbon and hydrogen fluoride.) If you are interested in the reaction with, say, chlorine, all you have to do is to replace Br by Cl in all the equations on this page. The mechanism for the reaction between propene and bromine The reaction is an example of electrophilic addition.
Warning! Just as with symmetrical alkenes, there are two versions of the propene / bromine mechanism in common use, and you must know which your examiners will accept. How you can find out which one your examiners expect is explained on the page on the addition of bromine to symmetrical alkenes.

Bromine as an electrophile

The bromine is a very "polarisable" molecule and the approaching pi bond in the propene induces a dipole in the bromine molecule. If you draw this mechanism in an exam, write the words "induced dipole" next to the bromine molecule. The simplified version of the mechanism
Note: Use this version unless your examiners insist on the more accurate one.

The more accurate version of the mechanism


Note: Don't learn this unless your examiners insist on it. There's no point in making life difficult for yourself.

In the first stage of the reaction, one of the bromine atoms becomes attached to both carbon atoms, with the positive charge being found on the bromine atom. A bromonium ion is formed.

The bromonium ion is then attacked from the back by a bromide ion formed in a nearby reaction.

EXPLAINING THE REACTION BETWEEN UNSYMMETRICAL ALKENES AND BROMINE


This page guides you through the mechanism for the electrophilic addition of bromine to unsymmetrical alkenes like propene.
Important! You will find it easier to make sense of this page if you first read about the electrophilic addition reactions between bromine and symmetrical alkenes like ethene, and addition to unsymmetrical alkenes in general. You may want to follow other links from those pages as well before you come back here again.

The electrophilic addition of bromine to propene


The attraction between the propene and the bromine The double bond in all alkenes is made up of two different parts. One pair of electrons lies on the line between the two nuclei where you would expect them to be. This is called a sigma bond. The other pair lies in an orbital above and below the plane of the rest of the molecule, and is called a pi bond. The pi bond is weaker than a sigma bond and is very vulnerable to attack.
Note: If this isn't fairly obvious to you, you really should follow the links at the top of the page before you go on - and perhaps explore other simpler reactions from the electrophilic addition menu as well.

As the bromine molecule approaches the pi bond, the electrons in that bond repel the electrons in the bromine-bromine bond down towards the bottom bromine. That produces an induced dipole in the bromine molecule.

Help! What is an "induced dipole"? A dipole is simply a separation of charge between + at one end and - at the other. "Induced" means that it has been created by some external influence (in this case the approach of the pi bond), and didn't already exist. Where it does already exist - as, for example, in HBr - it is called a permanent dipole.

The simplified version of the mechanism


Note: Use this version unless your examiners insist on the more accurate one. If you've come into this web site from a search engine directly to this page, read the notes on the addition of bromine to ethene before you go any further. Use the BACK button on your browser to return to this page.

The electrons from the pi bond move down towards the slightly positive bromine atom.

In the process, the electrons in the Br-Br bond are repelled down until they are entirely on the bottom bromine atom, producing a bromide ion. Notice the way that the pi bond electrons have moved. By swinging so that the bromine is attached to the right-hand carbon, a secondary carbocation has been formed. That is more stable than the primary one which would have been formed if the pi electrons had swung the other way.
Note: If this doesn't make sense to you, read about carbocations (previously called carbonium ions) and addition to unsymmetrical alkenes in general.

In the second stage of the mechanism, the lone pair of electrons on the bromide ion is strongly attracted to the positive carbon and moves towards it until a bond is formed.

The overall mechanism is therefore

The more accurate version of the mechanism

Note: Don't learn this unless you have to. There is a real risk of getting confused. If your examiners are happy to accept the simple version, there's no point in making life difficult for yourself.

The reaction starts off just the same as in the simplified version, with the pi bond electrons moving down towards the slightly positive bromine atom. But this time, the top bromine atom becomes attached to both carbon atoms, with the positive charge being found on the bromine rather than on one of the carbons. A bromonium ion is formed.

The bromonium ion is then attacked from the back by a bromide ion formed in a nearby reaction. It can't be attacked by its original bromide ion because the bromonium ion is completely cluttered up with a positive bromine on that side.

It doesn't matter which of the carbon atoms which were originally part of the double bond the bromide ion attacks - the end result would be just the same.
Note: You can't really draw this mechanism tidily in one line because the bromide ion has to be in a different place at the beginning of the second stage than it was at the end of the first stage.

CARBOCATIONS (or CARBONIUM IONS)


All carbocations (previously known as carbonium ions) carry a positive charge on a carbon atom. The name tells you that - a cation is a positive ion, and the "carbo" bit refers to a carbon atom. However there are important differences in the structures of various types of carbocations.

The different kinds of carbocations


Primary carbocations In a primary (1) carbocation, the carbon which carries the positive charge is only attached to one other alkyl group.
Help! An alkyl group is a group such as methyl, CH 3, or ethyl, CH3CH2. These are groups containing chains of carbon atoms which may be branched. Alkyl groups are given the general symbol R.

Some examples of primary carbocations include:

Notice that it doesn't matter how complicated the attached alkyl group is. All you are doing is counting the number of bonds from the positive carbon to other carbon atoms. In all the above cases there is only one such link. Using the symbol R for an alkyl group, a primary carbocation would be written as in the box.

Secondary carbocations In a secondary (2) carbocation, the carbon with the positive charge is attached to two other alkyl groups, which may be the

same or different. Examples:

A secondary carbocation has the general formula shown in the box. R and R' represent alkyl groups which may be the same or different.

Tertiary carbocations In a tertiary (3) carbocation, the positive carbon atom is attached to three alkyl groups, which may be any combination of same or different.

A tertiary carbocation has the general formula shown in the box. R, R' and R" are alkyl groups and may be the same or different.

The stability of the various carbocations


The "electron pushing effect" of alkyl groups You are probably familiar with the idea that bromine is more electronegative than hydrogen, so that in a H-Br bond the electrons are held closer to the bromine than the hydrogen. A bromine atom attached to a carbon atom would have precisely the same effect - the electrons being pulled towards the bromine end of the bond. The bromine has a negative inductive effect.
Help! If you aren't familiar with all of this, follow this link to read about electronegativity and bond polarity before you go any further.

Use the BACK button on your browser to return to this page.

Alkyl groups do precisely the opposite and, rather than draw electrons towards themselves, tend to "push" electrons away.
Note: The term "electron pushing" is only to help remember what happens. The alkyl group doesn't literally "push" the electrons away the other end of the bond attracts them more strongly.

This means that the alkyl group becomes slightly positive ( +) and the carbon they are attached to becomes slightly negative ( -). The alkyl group has a positive inductive effect. This is sometimes shown as, for example:

The arrow shows the electrons being "pushed" away from the CH3 group. The plus sign on the left-hand end of it shows that the CH3 group is becoming positive. The symbols + and simply reinforce that idea. The importance of spreading charge around in making ions stable The general rule-of-thumb is that if a charge is very localised (all concentrated on one atom) the ion is much less stable than if the charge is spread out over several atoms. Applying that to carbocations of various sorts . . .

You will see that the electron pushing effect of the CH3 group is placing more and more negative charge on the positive carbon

as you go from primary to secondary to tertiary carbocations. The effect of this, of course, is to cut down that positive charge. At the same time, the region around the various CH3 groups is becoming somewhat positive. The net effect, then, is that the positive charge is being spread out over more and more atoms as you go from primary to secondary to tertiary ions. The more you can spread the charge around, the more stable the ion becomes. Order of stability of carbocations primary < secondary < tertiary
Note: The symbol "<" means "is less than". So what this is saying is that primary ions are less stable than secondary ones which in turn are less stable than tertiary ones.

The stability of the carbocations in terms of energetics When we talk about secondary carbocations being more stable than primary ones, what exactly do we mean? We are actually talking about energetic stability - secondary carbocations are lower down an energy "ladder" than primary ones. This means that it is going to take more energy to make a primary carbocation than a secondary one. If there is a choice between making a secondary ion or a primary one, it will be much easier to make the secondary one. Similarly, if there is a choice between making a tertiary ion or a secondary one, it will be easier to make the tertiary one. This has important implications in the reactions of unsymmetrical alkenes. If you are interested in these, follow the link below to the electrophilic addition reactions menu.

ELECTROPHILIC ADDITION TO UNSYMMETRICAL ALKENES


Important! To make sense of this page, you will need to understand about the structure and stability of carbocations (previously called carbonium ions) and be confident about electrophilic addition to simple alkenes like ethene. If you aren't sure about either of these things, follow these links first. Trying to build on shaky foundations risks confusing you and undermining your confidence.

The Problem
The addition of H-X to an unsymmetrical alkene like propene An unsymmetrical alkene is one like propene or but-1-ene in which the groups or atoms attached to either end of the carboncarbon double bond are different. For example, in propene there are a hydrogen and a methyl group at one end, but two hydrogen atoms at the other end of the double bond.

With these unsymmetrical alkenes, it is possible to get two different products during some addition reactions. During the addition of a molecule HX to propene, you could in principle get either this reaction:

or this one:

It depends on which way around you add the HX across the double bond. In fact, in most cases, it's mainly the second reaction which happens. The hydrogen atom becomes attached to the righthand carbon atom as we've drawn it. Markovnikov's Rule
Note: Don't worry too much about the spelling of Markovnikov - there are nearly as many versions as there are text books. It's an English attempt at a Russian name and Markovnikov wouldn't actually have recognised any of them!

When a compound HX is added to an unsymmetrical alkene, the hydrogen becomes attached to the carbon with the most hydrogens attached to it already. Remember that the HX has to attach itself to the carbon atoms which were originally part of the double bond. So in this case, adding HX to CH3CH=CH2, the hydrogen is attached to the CH2 group, because the CH2 group has more hydrogens than the CH group. Notice that only the hydrogens directly attached count. The ones in the CH3 group are totally irrelevant.
Warning! Markovnikov's Rule is a useful guide for you to work out which way round to add something across a double bond, but it isn't the reason why things add that way. Propene has never even heard of Markovnikov! When the question arises in an exam, you will need a much more fundamental explanation which is coming up next. As a general principle, don't quote Markovnikov's Rule in an exam unless you are specifically asked for it.

The Mechanism

HX as an electrophile In each of the cases we are interested in, X is more electronegative than hydrogen. That means that the bonding pair of electrons will be dragged towards the X end of the bond, and so the hydrogen becomes slightly positively charged. The slightly positive hydrogen is the electrophile, and is attracted to the pi bond in the propene. What happens next determines which way around the HX adds across the double bond. The two possible mechanisms In both possibilities, the pi bond breaks and the electron pair swings down to form a bond with the hydrogen atom. At the same time, the electrons in the H-X bond are repelled right down on to the X to give an X- ion. The difference lies in which way the electron pair in the pi bond swings. First possibility The electron pair moves to form a bond between the hydrogen and the left-hand carbon.

When the second stage of the mechanism happens, and the lone pair on X- forms a bond with the positive carbon atom, the product of this mechanism is not the one which Markovnikov's Rule predicts. Second possibility The electron pair moves to form a bond between the hydrogen and the right-hand carbon.

This time, the overall mechanism leads to the correct product. Why does one of these work better than the other? That's the truth of the situation! One of these mechanisms works better than the other one. The second mechanism works much faster than the first, and so most of the product that you get is CH3-CHX-CH3. There will be small amounts of CH3-CH2-CH2X despite what Markovnikov says! The difference between the two mechanisms lies in the intermediates - the things formed at the half-way stage. In the mechanism that works best, you get a secondary carbocation formed as one of the intermediate ions.

In the slow mechanism which produces hardly any product, you get a primary carbocation formed instead.

It is much easier to form the secondary carbocation because it is more energetically stable. The activation energy for the reaction will be less, and so most of the reaction happens via that mechanism.
Activation energy: The minimum energy needed before a reaction will occur. In this case it is the energy needed to break the various bonds and make the carbocation and the X ion.

How to attack this sort of question in an exam Suppose you were asked for the mechanism for the addition of HX to some alkene you hadn't come across before. First, you need to decide whether the alkene is symmetrical or not. If it is symmetrical, there's no problem - it wouldn't matter which way around you added the HX. If it is unsymmetrical, you need to decide which way round the HX is going to add. For example, supposed you were asked for the mechanism for the addition of HX to but-1-ene, CH3-CH2-CH=CH2. First, use Markovnikov's Rule to decide which carbon to attach the hydrogen to. In this case, the hydrogen would get attached to the CH2 end of the double bond, because that carbon has more hydrogens than the CH end.
Warning! Markovnikov's Rule is only to help you decide. Don't give the examiners any hint that you are using it - unless they specifically ask.

Now write the mechanism, taking care to draw the curly arrow showing the movement of the pi bond so that the hydrogen gets attached to that particular CH2 carbon.

If you are asked why the HX adds this way round, look at the carbocation formed as an intermediate and decide whether it is secondary or tertiary. Here it is a secondary ion. Then think about what sort of ion would be formed if the HX added the other way around. In this case that would be a primary ion.

Then say something like: "The secondary carbocation formed in this reaction is more

energetically stable than the primary one which would be formed if the addition was the other way round, and so less activation energy is needed."
Important! If there were bits of this that you found you couldn't understand, you are probably trying to do too much too quickly. Follow the links to carbocations (carbonium ions) and electrophilic addition to simple alkenes, get the basics sorted out properly, and then try again.

THE REACTION BETWEEN UNSYMMETRICAL ALKENES AND THE HYDROGEN HALIDES


This page gives you the facts and a simple, uncluttered mechanism for the electrophilic addition reactions between the hydrogen halides and alkenes like propene. Hydrogen halides include hydrogen chloride and hydrogen bromide. If you want the mechanisms explained to you in detail, there is a link at the bottom of the page. An unsymmetrical alkene is one like propene in which the groups or atoms attached to either end of the carbon-carbon double bond are different. For example, in propene there are a hydrogen and a methyl group at one end, but two hydrogen atoms at the other end of the double bond. But-1-ene is another unsymmetrical alkene.

Electrophilic addition reactions involving hydrogen bromide

The facts As with all alkenes, unsymmetrical alkenes like propene react with hydrogen bromide in the cold. The double bond breaks and a hydrogen atom ends up attached to one of the carbons and a bromine atom to the other. In the case of propene, 2-bromopropane is formed.

This would normally be written in a more condensed form as

The product is 2-bromopropane.


Note: There is another possible reaction between unsymmetrical alkenes and hydrogen bromide (but not the other hydrogen halides) unless the hydrogen bromide and alkene are absolutely pure. A different mechanism happens (a free radical chain reaction - not on UK A' level syllabuses) which leads to the hydrogen and bromine adding the opposite way round. For A' level purposes, you don't need to worry about that. However, if you are interested, you will find the free radical addition mechanism by following this link. Use the BACK button on your browser to return to this page later.

This is in line with Markovnikov's Rule which says: When a compound HX is added to an unsymmetrical alkene, the hydrogen becomes attached to the carbon with the most hydrogens attached to it already. In this case, the hydrogen becomes attached to the CH2 group, because the CH2 group has more hydrogens than the CH group. Notice that only the hydrogens directly attached to the carbon atoms at either end of the double bond count. The ones in the

CH3 group are totally irrelevant.


Warning! Markovnikov's Rule is a useful guide for you to work out which way round to add something across a double bond, but it isn't the reason why things add that way. As a general principle, don't quote Markovnikov's Rule in an exam unless you are specifically asked for it.

The mechanism This is an example of electrophilic addition.

The addition is this way around because the intermediate carbocation (previously called a carbonium ion) formed is secondary. This is more stable (and so is easier to form) than the primary carbocation which would be produced if the hydrogen became attached to the centre carbon atom and the bromine to the end one.

Electrophilic addition reactions involving the other hydrogen halides


The facts Hydrogen fluoride, hydrogen chloride and hydrogen iodide all add on in exactly the same way as hydrogen bromide. The only differences lie in the rates of reaction: HF HCl HBr HI fastest reaction slowest reaction

This is because the hydrogen-halogen bond gets weaker as the halogen atom gets bigger. If the bond is weaker, it breaks more easily and so the reaction is faster. If the halogen is given the symbol X, the equation for the reaction with propene is:

Notice that the product is still in line with Markovnikov's Rule. The mechanism These are still examples of electrophilic addition. Again using X to stand for any halogen:

Again, the intermediate carbocation formed is secondary. This is more stable than the primary carbocation ion which would be formed if the hydrogen attached to the centre carbon atom and the X to the end one. If it is more stable it will be easier to make.

EXPLAINING THE REACTION BETWEEN UNSYMMETRICAL ALKENES AND THE HYDROGEN HALIDES
This page guides you through the mechanism for the electrophilic addition of hydrogen halides such as hydrogen bromide to unsymmetrical alkenes like propene.
Important! To make sense of this page, you will need to

understand about the structure and stability of carbocations (previously called carbonium ions) and be confident about electrophilic addition to simple alkenes like ethene. If you aren't sure about either of these things, follow these links first. You would also find it easier if you first read about the electrophilic addition reactions between hydrogen halides and symmetrical alkenes like ethene, and addition to unsymmetrical alkenes in general. If you have just come from those, ignore these links!

Electrophilic addition reactions involving hydrogen bromide


If you want the mechanism for one of the other hydrogen halides, simply replace Br by whatever else you are interested in - F or Cl or I. There is no difference whatsoever in the mechanisms. You might, however, need to be aware that there is an alternative mechanism involving hydrogen bromide and alkenes if the reaction mixture is impure in the presence of organic peroxides or oxygen from the air.
Note: The different mechanism (a free radical chain reaction - not on UK A' level syllabuses) leads to the hydrogen and bromine adding the opposite way round. For A' level purposes, you don't need to worry about that. However, if you are interested, you will find the free radical addition mechanism by following this link. Use the BACK button on your browser to return to this page later.

Hydrogen bromide as an electrophile Hydrogen bromide is chosen as a typical hydrogen halide. Bromine is more electronegative than hydrogen. That means that the bonding pair of electrons is pulled towards the bromine end of the bond, and so the hydrogen bromide molecule is polar.

Note: If you aren't sure about electronegativity and bond polarity follow this link before you read on. Use the BACK button on your browser to return to this page.

The slightly positive hydrogen atom will be attracted to negative regions in other molecules, and is therefore an electrophile.
Electrophile: A substance with a strong attraction to a negative region in another substance. Electrophiles are either fully positive ions, or the slightly positive end of a polar molecule.

The reaction of propene with hydrogen bromide The double bond in all alkenes is made up of two different parts. One pair of electrons lies on the line between the two nuclei where you would expect them to be. This is called a sigma bond. The other pair lies in an orbital above and below the plane of the rest of the molecule, and is called a pi bond. The pi bond is weaker than a sigma bond and is very vulnerable to attack.
Note: If this isn't fairly obvious to you, you really ought to read the page introducing electrophilic addition before you go on. Use the BACK button on your browser to return to this page.

As the HBr approaches the pi bond, the electrons in that bond are attracted towards the slightly positive hydrogen atom. That repels the electrons in the hydrogen-bromine bond down towards the bromine.

The electron movements continue until a new bond is made between one of the carbon atoms and the hydrogen. The bromine now has both electrons from the H-Br bond, and so is negatively charged as a bromide ion. The problem is that there are two possible ways that the pi bond electrons could move. They could form a bond between the hydrogen and the left-hand carbon:

or they could form a bond with the right-hand one:

It's the second of these changes that happens more readily. In that case, a secondary carbocation is formed - and that's more energetically stable than the primary one formed in the first possibility. Because the secondary ion is more energetically stable, it will form more easily and so the reaction needs less activation

energy.
Important! If you don't understand about the structure and stability of carbocations (carbonium ions) follow this link. Activation energy: The minimum energy needed before a reaction will occur. In this case it is the energy needed to break the various bonds and make the carbocation and the bromide ion.

Once the ions have been formed, the lone pair on the bromide ion is strongly attracted towards the positive carbon atom. It moves towards it and forms a bond.

Note: There are actually 4 lone pairs around the bromide ion, but we are only interested in the one shown.

That leaves you with the over-all mechanism:

Important! If you've had problems with this page you might find it useful to read about addition to unsymmetrical alkenes in general, and then come back here again.

THE REACTION BETWEEN UNSYMMETRICAL ALKENES AND SULPHURIC ACID

This page gives you the facts and a simple, uncluttered mechanism for the electrophilic addition reactions between sulphuric acid and unsymmetrical alkenes like propene. If you want the mechanism explained to you in detail, there is a link at the bottom of the page.

The electrophilic addition reaction between propene and sulphuric acid


The facts Alkenes react with concentrated sulphuric acid in the cold to produce alkyl hydrogensulphates. In the case of propene, the equation is:

Note: If you are confused about the structure of the product, a similar structure is described in more detail in the page about the reaction between sulphuric acid and symmetrical alkenes.

This is in line with Markovnikov's Rule which says: When a compound HX is added to an unsymmetrical alkene, the hydrogen becomes attached to the carbon with the most hydrogens attached to it already. In this case, the hydrogen becomes attached to the CH2 group, because the CH2 group has more hydrogens than the CH group. Notice that only the hydrogens directly attached to the carbon atoms at either end of the double bond count. The ones in the CH3 group are totally irrelevant. The mechanism This is an example of electrophilic addition.

The addition is this way around because it is easier to produce the secondary carbocation (carbonium ion) than the primary one which would be formed if the hydrogen became attached to the centre carbon atom and the rest of the sulphuric acid to the end one.

EXPLAINING THE REACTION BETWEEN UNSYMMETRICAL ALKENES AND SULPHURIC ACID

This page guides you through the mechanism for the electrophilic addition of sulphuric acid to unsymmetrical alkenes like propene. The electrophilic addition reaction between propene and sulphuric acid
Important! The reaction between propene and sulphuric acid has exactly the same mechanism as the one involving hydrogen bromide. It just looks more complicated because of the structure of sulphuric acid. Make sure that you understand the mechanism for the reaction between propene and HBr before you read on. It is also essential that you read about the reaction between ethene and sulphuric acid before you worry about the additional problem caused by an unsymmetrical alkene like propene.

The structure of sulphuric acid Compare the structure of sulphuric acid with that of hydrogen

bromide:

We are focussing on only one of the hydrogens in the sulphuric acid because the other one will be pointing away from the double bond in the alkene as the molecules approach each other. In each case, the hydrogen is attached to a more electronegative element, and so carries a slight positive charge. That means that the hydrogen atoms will serve as electrophiles. When the sulphuric acid reacts, the whole of the shaded part of the molecule remains as a complete unit. What happens to that unit is exactly the same as happens to the bromine in the reaction involving HBr. The mechanism Remember that the active part of the double bond is the pi bond which lies in an orbital above and below the plane of the rest of the molecule. The pi bond is very vulnerable to attack.
Note: If this isn't fairly obvious to you, you really ought to read the page introducing electrophilic addition before you go on. Use the BACK button on your browser to return to this page.

The slightly positive hydrogen atom in the sulphuric acid acts as an electrophile, and is strongly attracted to the electrons in the pi bond. As the sulphuric acid approaches the pi bond, the electrons in that bond are drawn down towards the slightly positive hydrogen atom. That repels the electrons in the hydrogen-oxygen bond down towards the oxygen.

The electron movements continue until a new bond is made between one of the carbon atoms and the hydrogen. The oxygen now has both electrons from the H-O bond, and so becomes negatively charged. The problem is that there are two possible ways that the pi bond electrons could move. They could form a bond between the hydrogen and the left-hand carbon:

or they could form a bond with the right-hand one:

It's the second of these changes that happens more readily. In that case, a secondary carbocation is formed - and that's more energetically stable than the primary one formed in the first

possibility. Because the secondary ion is more energetically stable, it will form more easily and so the reaction needs less activation energy - and so happens faster.
Important! If you don't understand about the structure and stability of carbocations (previously known as carbonium ions) follow this link.

Once the ions have been formed, the lone pair on the negative oxygen is strongly attracted towards the positive carbon atom. It moves towards it and forms a bond.

That leaves you with the over-all mechanism:

THE REACTION BETWEEN UNSYMMETRICAL ALKENES AND BROMINE


Important! This page assumes that you have already read the page on the addition of bromine to symmetrical alkenes. If you haven't, you must read it before you go on. It contains important advice that you will need to make best use of this page.

This page gives you the facts and a simple, uncluttered

mechanism for the electrophilic addition reactions between bromine and alkenes like propene. If you want the mechanism explained to you in detail, there is a link at the bottom of the page. An unsymmetrical alkene is one like propene in which the groups or atoms attached to either end of the carbon-carbon double bond are different. For example, in propene there are a hydrogen and a methyl group at one end, but two hydrogen atoms at the other end of the double bond. But-1-ene is another unsymmetrical alkene.

The electrophilic addition of bromine to propene


The facts In common with all other alkenes, propene reacts in the cold with pure liquid bromine, or with a solution of bromine in an organic solvent like tetrachloromethane. The double bond breaks, and a bromine atom becomes attached to each carbon. The bromine loses its original red-brown colour to give a colourless liquid. In the case of the reaction with propene, 1,2dibromopropane is formed.

The other halogens, apart from fluorine, behave similarly. (Fluorine reacts explosively with all hydrocarbons - including alkenes - to give carbon and hydrogen fluoride.) If you are interested in the reaction with, say, chlorine, all you have to do is to replace Br by Cl in all the equations on this page. The mechanism for the reaction between propene and bromine

The reaction is an example of electrophilic addition.


Warning! Just as with symmetrical alkenes, there are two versions of the propene / bromine mechanism in common use, and you must know which your examiners will accept. How you can find out which one your examiners expect is explained on the page on the addition of bromine to symmetrical alkenes.

Bromine as an electrophile The bromine is a very "polarisable" molecule and the approaching pi bond in the propene induces a dipole in the bromine molecule. If you draw this mechanism in an exam, write the words "induced dipole" next to the bromine molecule. The simplified version of the mechanism
Note: Use this version unless your examiners insist on the more accurate one.

The more accurate version of the mechanism


Note: Don't learn this unless your examiners insist on it. There's no point in making life difficult for yourself.

In the first stage of the reaction, one of the bromine atoms becomes attached to both carbon atoms, with the positive charge being found on the bromine atom. A bromonium ion is formed.

The bromonium ion is then attacked from the back by a bromide ion formed in a nearby reaction.

EXPLAINING THE REACTION BETWEEN UNSYMMETRICAL ALKENES AND BROMINE


This page guides you through the mechanism for the electrophilic addition of bromine to unsymmetrical alkenes like propene.
Important! You will find it easier to make sense of this page if you first read about the electrophilic addition reactions between bromine and symmetrical alkenes like ethene, and addition to unsymmetrical alkenes in general. You may want to follow other links from those pages as well before you come back here again.

The electrophilic addition of bromine to propene


The attraction between the propene and the bromine The double bond in all alkenes is made up of two different parts. One pair of electrons lies on the line between the two nuclei where you would expect them to be. This is called a sigma bond. The other pair lies in an orbital above and below the plane of the rest of the molecule, and is called a pi bond. The pi bond is

weaker than a sigma bond and is very vulnerable to attack.


Note: If this isn't fairly obvious to you, you really should follow the links at the top of the page before you go on - and perhaps explore other simpler reactions from the electrophilic addition menu as well.

As the bromine molecule approaches the pi bond, the electrons in that bond repel the electrons in the bromine-bromine bond down towards the bottom bromine. That produces an induced dipole in the bromine molecule.

Help! What is an "induced dipole"? A dipole is simply a separation of charge between + at one end and - at the other. "Induced" means that it has been created by some external influence (in this case the approach of the pi bond), and didn't already exist. Where it does already exist - as, for example, in HBr - it is called a permanent dipole.

The simplified version of the mechanism


Note: Use this version unless your examiners insist on the more accurate one. If you've come into this web site from a search engine directly to this page, read the notes on the addition of bromine to ethene before you go any further. Use the BACK button on your browser to return to this page.

The electrons from the pi bond move down towards the slightly

positive bromine atom.

In the process, the electrons in the Br-Br bond are repelled down until they are entirely on the bottom bromine atom, producing a bromide ion. Notice the way that the pi bond electrons have moved. By swinging so that the bromine is attached to the right-hand carbon, a secondary carbocation has been formed. That is more stable than the primary one which would have been formed if the pi electrons had swung the other way.
Note: If this doesn't make sense to you, read about carbocations (previously called carbonium ions) and addition to unsymmetrical alkenes in general.

In the second stage of the mechanism, the lone pair of electrons on the bromide ion is strongly attracted to the positive carbon and moves towards it until a bond is formed.

The overall mechanism is therefore

The more accurate version of the mechanism

Note: Don't learn this unless you have to. There is a real risk of getting confused. If your examiners are happy to accept the simple version, there's no point in making life difficult for yourself.

The reaction starts off just the same as in the simplified version, with the pi bond electrons moving down towards the slightly positive bromine atom. But this time, the top bromine atom becomes attached to both carbon atoms, with the positive charge being found on the bromine rather than on one of the carbons. A bromonium ion is formed.

The bromonium ion is then attacked from the back by a bromide ion formed in a nearby reaction. It can't be attacked by its original bromide ion because the bromonium ion is completely cluttered up with a positive bromine on that side.

It doesn't matter which of the carbon atoms which were originally part of the double bond the bromide ion attacks - the end result would be just the same.
Note: You can't really draw this mechanism tidily in one line because the bromide ion has to be in a different place at the beginning of the second stage than it was at the end of the first stage.

ELECTROPHILIC SUBSTITUTION MECHANISMS MENU

What is electrophilic substitution? . . . Some background on benzene (including links to more detailed discussions) and a general mechanism which covers several of benzene's reactions. The nitration of benzene . . . The mechanism for the formation of nitrobenzene from benzene. The Friedel-Crafts acylation of benzene . . . The mechanism for the substitution of an acyl group such as CH3CO into benzene. The Friedel-Crafts alkylation of benzene . . . The mechanism for the substitution of an alkyl group such as CH3 into benzene. An industrial alkylation of benzene . . . The mechanism for the substitution of an alkyl group such as CH3CH2 into benzene, by a reaction involving an alkene such as ethene. The halogenation of benzene . . . The mechanism for the substitution of atoms like chlorine and bromine into benzene rings. The sulphonation of benzene . . . The mechanism for reaction between benzene and concentrated sulphuric acid to produce benzenesulphonic acid Some substitution reactions of methylbenzene . . . Illustrates how to cope with the problem of substituting things into rings which already have something else attached.

ELECTROPHILIC SUBSTITUTION
Background

Electrophilic substitution happens in many of the reactions of compounds containing benzene rings - the arenes. For simplicity, we'll only look for now at benzene itself.
Note: Before you start it would be a good idea if you had a clear idea about the structure of benzene. Check your syllabus now to find out what you need to know, and then read the page on the modern orbital view of benzene in the organic bonding section of this site. Don't forget to look in the section(s) in your syllabus on bonding as well as organic chemistry. Haven't got a syllabus? If you are working towards a UKbased exam (A level or its equivalent), follow this link to find out how to get one.

This is what you need to understand for the purposes of the electrophilic substitution mechanisms:

Benzene, C6H6, is a planar molecule containing a ring of six carbon atoms each with a hydrogen atom attached. There are delocalised electrons above and below the plane of the ring.

The presence of the delocalised electrons makes benzene particularly stable. Benzene resists addition reactions because that would involve breaking the delocalisation and losing that stability.

Benzene is represented by this symbol, where the circle represents the delocalised electrons, and each corner of the hexagon has a carbon atom with a hydrogen attached.

Electrophilic substitution reactions involving positive ions


Benzene and electrophiles Because of the delocalised electrons exposed above and below the plane of the rest of the molecule, benzene is obviously going to be highly attractive to electrophiles - species which seek after electron rich areas in other molecules.
Species: A useful word which can mean any particle you want it to mean - an atom, a molecule, an ion or a free radical.

The electrophile will either be a positive ion, or the slightly positive end of a polar molecule.
Help! If you aren't sure what a polar molecule is, read about electronegativity and polar bonds before you go on. Use the BACK button on your browser to return to this page.

The delocalised electrons above and below the plane of the benzene molecule are open to attack in the same way as those above and below the plane of an ethene molecule. However, the end result will be different.

If benzene underwent addition reactions in the same way as ethene, it would need to use some of the delocalised electrons to form bonds with the new atoms or groups. This would break the delocalisation - and this costs energy.

Note: You can read about electrophilic addition to ethene if you are interested. Use the BACK button on your browser to return to this page.

Instead, it can maintain the delocalisation if it replaces a hydrogen atom by something else - a substitution reaction. The hydrogen atoms aren't involved in any way with the delocalised electrons. In most of benzene's reactions, the electrophile is a positive ion, and these reactions all follow a general pattern. The general mechanism The first stage Suppose the electrophile is a positive ion X+. Two of the electrons in the delocalised system are attracted towards the X+ and form a bond with it. This has the effect of breaking the delocalisation, although not completely.

Note: If you aren't sure about the use of curly arrows in mechanisms, you must follow this link before you go on. Use the BACK button on your browser to return to this page.

The ion formed in this step isn't the final product. It immediately goes on to react with something else. It is just an intermediate. There is still delocalisation in the intermediate formed, but it only covers part of the ion. When you write one of these mechanisms, draw the partial delocalisation to take in all the carbon atoms apart from the one that the X has become attached to. The intermediate ion carries a positive charge because you are joining together a neutral molecule and a positive ion. This positive charge is spread over the delocalised part of the ring. Simply draw the "+" in the middle of the ring. The hydrogen at the top isn't new - it's the hydrogen that was already attached to that carbon. We need to show that it is there for the next stage. The second stage

Here we've introduced a new ion, Y-. Where did this come from? You have to remember that it is impossible to get a positive ion on its own in a chemical system - so Y- is simply the negative ion that was originally associated with X+. Don't worry about this at the moment - it's much easier to see when you've got a real example in front of you. A lone pair of electrons on Y- forms a bond with the hydrogen atom at the top of the ring. That means that the pair of electrons joining the hydrogen onto the ring aren't needed any more. These then move down to plug the gap in the delocalised electrons, so restoring the delocalised ring of electrons which originally gave the benzene its special stability. The energetics of the reaction

The complete delocalisation is temporarily broken as X replaces H on the ring, and this costs energy. However, that energy is recovered when the delocalisation is re-established. This initial input of energy is simply the activation energy for the reaction. In this case, it is going to be high (something around 150 kJ mol-1), and this means that benzene's reactions tend to be slow.

Electrophilic substitution reactions not involving positive ions


Halogenation and sulphonation In these reactions, the electrophiles are polar molecules rather than fully positive ions. Because these mechanisms are different from what's gone before (and from each other), there isn't any point in dealing with them in a general way. You will find them explained in full if you follow the link to the electrophilic substitution menu below.

THE NITRATION OF BENZENE


This page gives you the facts and a simple, uncluttered mechanism for the electrophilic substitution reaction between benzene and a mixture of concentrated nitric acid and concentrated sulphuric acid. If you want the nitration mechanism explained to you in detail, there is a link at the bottom of the page.

The electrophilic substitution reaction between benzene and nitric acid


The facts Benzene is treated with a mixture of concentrated nitric acid and concentrated sulphuric acid at a temperature not exceeding 50C. As temperature increases there is a greater chance of getting more than one nitro group, -NO2, substituted onto the ring. Nitrobenzene is formed.

or:

The concentrated sulphuric acid is acting as a catalyst.

The formation of the electrophile The electrophile is the "nitronium ion" or the "nitryl cation", NO2+. This is formed by reaction between the nitric acid and the sulphuric acid.

The electrophilic substitution mechanism Stage one

Stage two

EXPLAINING THE NITRATION OF BENZENE


This page guides you through the mechanism for the nitration of benzene involving an electrophilic substitution reaction between benzene and nitric acid.

Important! It would help if you first read the page What is electrophilic substitution? before you went on.

The electrophilic substitution reaction between benzene and nitric acid


The formation of the electrophile If you are going to substitute an -NO2 group into the ring, then the electrophile must be NO2+. This is called the "nitronium ion" or the "nitryl cation", and is formed by reaction between the nitric acid and sulphuric acid.
Note: If you don't understand why the electrophile + has got to be NO2 , then you really should look at What is electrophilic substitution? before you go on. If you are going to substitute X onto the ring, then the + electrophile must be X . If you are going to insert an NO2 group onto the ring, then the electrophile must + be NO2 .

The equation

Important! Check past exam papers to see whether you will need to quote this equation in an exam. It's probable that you will, so learn it! There are ways of building it up, but they involve much more effort than learning it in the first place.

The hydrogensulphate ion, HSO4-, will also be involved in the mechanism. The hydroxonium ion, H3O+, isn't involved.
Note: The hydroxonium ion (also called the hydronium ion or oxonium ion) is simply a hydrogen ion attached to a water molecule - what is often + written more simply as H (aq).

The electrophilic substitution mechanism Stage one As the NO2+ ion approaches the delocalised electrons in the benzene, those electrons are strongly attracted towards the positive charge. Two electrons from the delocalised system are used to form a new bond with the NO2+ ion. Because those two electrons aren't a part of the delocalised system any longer, the delocalisation is partly broken, and in the process the ring gains a positive charge.

The hydrogen shown on the ring is the one which was already attached to that top carbon atom - it's nothing new or subtle! We need to show it there because it has to be removed in the second stage. Stage two The second stage involves a hydrogensulphate ion, HSO4-, which was produced at the same time as the NO2+ ion (refer back to the equation showing the formation of the electrophile if you've forgotten).
Note: Only one of the lone pairs in the hydrogensulphate ion is shown. There are lots more, but those aren't involved in the reaction.

This removes a hydrogen from the ring to form sulphuric acid the catalyst has therefore been regenerated. The electrons which originally joined the hydrogen to the ring are now used to re-establish the delocalised system. Stage two - a sloppy way of writing the same thing! You will often find the second stage of this reaction simplified in many (or even most!) books. The second stage is shown as:

The hydrogen is shown as "falling off" the ring as a hydrogen ion. This is sloppy and unsatisfactory on two counts:

Hydrogen ions never exist on their own in a chemical reaction. A hydrogen ion is a raw proton - the most intensely positive thing you can imagine. It will always be attached to something else. By not showing the hydrogensulphate ion, you can't show that the sulphuric acid catalyst has been regenerated. That's simply unsatisfying!

Showing exactly how the hydrogen is removed from the ring isn't difficult - do it properly!

THE FRIEDEL-CRAFTS ACYLATION OF BENZENE

This page gives you the facts and a simple, uncluttered mechanism for the electrophilic substitution reaction between benzene and ethanoyl chloride in the presence of an aluminium chloride catalyst. If you want the Friedel-Crafts acylation

mechanism explained to you in detail, there is a link at the bottom of the page.

The electrophilic substitution reaction between benzene and ethanoyl chloride What is acylation? An acyl group is an alkyl group attached to a carbon-oxygen double bond. If "R" represents any alkyl group, then an acyl group has the formula RCO-. Acylation means substituting an acyl group into something - in this case, into a benzene ring. The most commonly used acyl group is CH3CO-. This is called the ethanoyl group. In the example which follows we are substituting a CH3CO- group into the ring, but you could equally well use any other alkyl group instead of the CH3. The facts The most reactive substance containing an acyl group is an acyl chloride (also known as an acid chloride). These have the general formula RCOCl. Benzene is treated with a mixture of ethanoyl chloride, CH3COCl, and aluminium chloride as the catalyst. A ketone called phenylethanone is formed.
Note: Ketones: A family of compounds containing a carbon-oxygen double bond with a hydrocarbon group either side of it. In this case there is a methyl group on one side and a benzene ring on the other. Don't worry too much about the name "phenylethanone" - all that matters is that you can draw the structure.

or better:

The aluminium chloride isn't written into these equations because it is acting as a catalyst. If you wanted to include it, you could write AlCl3 over the top of the arrow.

The formation of the electrophile The electrophile is CH3CO+. It is formed by reaction between the ethanoyl chloride and the aluminium chloride catalyst.

The electrophilic substitution mechanism Stage one

Stage two

The hydrogen is removed by the AlCl4- ion which was formed at the same time as the CH3CO+ electrophile. The aluminium chloride catalyst is re-generated in this second stage.

EXPLAINING THE FRIEDEL-CRAFTS ACYLATION OF BENZENE

This page guides you through the mechanism for the FriedelCrafts acyation of benzene involving an electrophilic substitution reaction between benzene and ethanoyl chloride in the presence of an aluminium chloride catalyst.
Important! It would help if you first read the page What is electrophilic substitution? before you went on.

The electrophilic substitution reaction between benzene and ethanoyl chloride


The formation of the electrophile If you are going to replace a hydrogen atom in a benzene ring by CH3CO, then the electrophile must be the ion CH3CO+. The positive charge must be on the carbon atom, because that's what gets attached to the ring.
Note: If you don't understand why the electrophile has + got to be CH3CO , then you really should look at What is electrophilic substitution? before you go on. If you are going to substitute X onto the ring, then the + electrophile must be X . If you are going to insert an CH3CO group onto the ring, then the electrophile must be + CH3CO .

Aluminium chloride, AlCl3, is an electron deficient molecule. It is covalently bonded, but because the aluminium is only forming 3 bonds, and has no lone pairs, there are only 6 electrons around the aluminium atom rather than 8. It takes a chlorine (as a chloride ion) from the ethanoyl chloride, and forms a co-ordinate (dative covalent) bond with it.

Help! A co-ordinate bond is a covalent bond in which both electrons originally came from the same atom. In this case, both the electrons in the bond come from the chlorine being removed from the ethanoyl chloride.

The equation simplified

The electrophilic substitution mechanism Stage one As the CH3CO+ ion approaches the delocalised electrons in the benzene, those electrons are strongly attracted towards the positive charge. Two electrons from the delocalised system are used to form a new bond with the CH3CO+ ion. Because those two electrons aren't a part of the delocalised system any longer, the delocalisation is partly broken, and in the process the ring gains a positive charge.

Help! Don't be confused that we have reversed the way + + we have written the CH3CO ion to show it as COCH3. This is just so that it is easier to draw the mechanism tidily in the same way that the general mechanism was drawn on the page What is electrophilic substitution?

The hydrogen shown on the ring is the one which was already attached to that top carbon atom. We need to show it there because it has to be removed in the second stage. Stage two The second stage involves the AlCl4-, which was produced at the same time as the CH3CO+ ion.

One of the aluminium-chlorine bonds breaks and both electrons from it are used to join to the hydrogen. This removes the hydrogen from the ring to form HCl, and re-generates the aluminium chloride catalyst in the process. The electrons which originally joined the hydrogen to the ring are now used to reestablish the delocalised system.

THE FRIEDEL-CRAFTS ALKYLATION OF BENZENE

This page gives you the facts and a simple, uncluttered mechanism for the electrophilic substitution reaction between benzene and chloromethane in the presence of an aluminium chloride catalyst. Any other chloroalkane would work similarly. If you want the Friedel-Crafts alkylation mechanism explained to you in detail, there is a link at the bottom of the page.

The electrophilic substitution reaction between benzene and chloromethane What is alkylation? Alkylation means substituting an alkyl group into something - in this case into a benzene ring. A hydrogen on the ring is replaced

by a group like methyl or ethyl and so on. The facts Benzene is treated with a chloroalkane (for example, chloromethane or chloroethane) in the presence of aluminium chloride as a catalyst. On this page, we will look at substituting a methyl group, but any other alkyl group could be used in the same way. Substituting a methyl group gives methylbenzene - once known as toluene.

or better:

The aluminium chloride isn't written into these equations because it is acting as a catalyst. If you wanted to include it, you could write AlCl3 over the top of the arrow.
Note: The methylbenzene formed is more reactive than the original benzene, and so the reaction doesn't stop there. You get further methyl groups substituted around the ring. You won't have to worry about this for A' level.

The formation of the electrophile The electrophile is CH3+. It is formed by reaction between the chloromethane and the aluminium chloride catalyst.

The electrophilic substitution mechanism Stage one

Stage two

The hydrogen is removed by the AlCl4- ion which was formed at the same time as the CH3+ electrophile. The aluminium chloride catalyst is re-generated in this second stage.

EXPLAINING THE FRIEDEL-CRAFTS ALKYLATION OF BENZENE

This page guides you through the mechanism for the Friedel-Crafts alkyation of benzene involving an electrophilic substitution reaction between benzene and a chloroalkane like chloromethane in the presence of an aluminium chloride catalyst.
Important! It would help if you first read the page What is electrophilic substitution? before you went on.

The electrophilic substitution reaction between benzene and chloromethane


The formation of the electrophile

If you are going to replace a hydrogen atom in a benzene ring by CH3, then the electrophile must be the ion CH3+.
Note: If you don't understand why the electrophile + has got to be CH3 , then you really should look at What is electrophilic substitution? before you go on. If you are going to substitute X onto the ring, then + the electrophile must be X . If you are going to insert an CH3 group onto the ring, then the + electrophile must be CH3 .

Aluminium chloride, AlCl3, is an electron deficient molecule. It is covalently bonded, but because the aluminium is only forming 3 bonds, and has no lone pairs, there are only 6 electrons around the aluminium atom rather than 8. It takes a chlorine (as a chloride ion) from the chloromethane, and forms a co-ordinate (dative covalent) bond with it.

Help! A co-ordinate bond is a covalent bond in which both electrons originally came from the same atom. In this case, both the electrons in the bond come from the chlorine being removed from the chloromethane.

The equation simplified

The electrophilic substitution mechanism Stage one As the CH3+ ion approaches the delocalised electrons in the benzene, those electrons are strongly attracted towards the

positive charge. Two electrons from the delocalised system are used to form a new bond with the CH3+ ion. Because those two electrons aren't a part of the delocalised system any longer, the delocalisation is partly broken, and in the process the ring gains a positive charge.

The hydrogen shown on the ring is the one which was already attached to that top carbon atom. We need to show it there because it has to be removed in the second stage. Stage two The second stage involves the AlCl4-, which was produced at the same time as the CH3+ ion.

One of the aluminium-chlorine bonds breaks and both electrons from it are used to join to the hydrogen. This removes the hydrogen from the ring to form HCl, and re-generates the aluminium chloride catalyst in the process. The electrons which originally joined the hydrogen to the ring are now used to reestablish the delocalised system.

THE HALOGENATION OF BENZENE


This page gives you the facts and a simple, uncluttered mechanism for the electrophilic substitution reaction between benzene and chlorine or bromine in the presence of a catalyst such as aluminium chloride or iron. If you want this mechanism explained to you in detail, there is a link at the bottom of the page.

The electrophilic substitution reaction between benzene and chlorine or bromine


The facts Benzene reacts with chlorine or bromine in an electrophilic substitution reaction, but only in the presence of a catalyst. The catalyst is either aluminium chloride (or aluminium bromide if you are reacting benzene with bromine) or iron. Strictly speaking iron isn't a catalyst, because it gets permanently changed during the reaction. It reacts with some of the chlorine or bromine to form iron(III) chloride, FeCl 3, or iron(III) bromide, FeBr3.

These compounds act as the catalyst and behave exactly like aluminium chloride in these reactions. The reaction with chlorine The reaction between benzene and chlorine in the presence of either aluminium chloride or iron gives chlorobenzene.

or:

The reaction with bromine The reaction between benzene and bromine in the presence of either aluminium bromide or iron gives bromobenzene. Iron is usually used because it is cheaper and more readily available.

or:

The formation of the electrophile We are going to explore the reaction using chlorine and aluminium chloride. If you want one of the other combinations, all you have to do is to replace each Cl by Br, or each Al by Fe. As a chlorine molecule approaches the benzene ring, the delocalised electrons in the ring repel electrons in the chlorine-chlorine bond.

It is the slightly positive end of the chlorine molecule which acts as the electrophile. The presence of the aluminium chloride helps this polarisation. The electrophilic substitution mechanism Stage one

Stage two

The hydrogen is removed by the AlCl4- ion which was formed in the first stage. The aluminium chloride catalyst is re-generated in this second stage.

EXPLAINING THE HALOGENATION OF BENZENE


This page guides you through the mechanism for the electrophilic substitution reaction between benzene and chlorine in the presence of an aluminium chloride or an iron catalyst. The reaction involving bromine is exactly the same, except that iron would be the preferred catalyst. Aluminium bromide could be used as an alternative. In what follows, if you want one of the other combinations, all you have to do is to replace each Cl by Br, or each Al by Fe.
Note: The reason that iron functions in the same way as the aluminium compounds is explained in the "facts" section of the introductory page on halogenation of benzene. If you've forgotten, you might like to look back at that before you go on.

The electrophilic substitution reaction between benzene and chlorine


The formation of the electrophile Many of the electrophilic substitution reactions of benzene involve an attack on the benzene by a positive ion. In the chlorine case, forming a Cl+ ion needs too much energy. As the chlorine molecule approaches a benzene ring, the delocalised electrons in the ring repel the electrons in the chlorine-chlorine bond. That induces a dipole in the chlorine.
Note: If you aren't happy about the structure of benzene, you could follow this link. The formation of the induced dipole is much the same as happens in the addition of bromine to ethene. If you aren't sure about induced dipoles, you might like to have a look at the beginning of that page.

Also nearby is the aluminium chloride, and this encourages the polarisation of the chlorine. The aluminium chloride is an electron deficient molecule, with the aluminium only having 3 pairs of electrons in its bonding level. The aluminium is strongly attracted to the slightly negative end of the chlorine molecule, and pulls electrons even more towards that end. The electrophilic substitution mechanism Stage one Two electrons from the delocalised system are used to form a new bond with the slightly positive chlorine atom. Because those two electrons aren't a part of the delocalised system any longer, the delocalisation is partly broken, and in the process the ring gains a positive charge.

The hydrogen shown on the ring is the one which was already attached to that top carbon atom. We need to show it there because it has to be removed in the second stage. Notice that the chlorine-chlorine bond breaks, transferring a chloride ion to the AlCl3 to make an AlCl4- ion. Stage two The second stage involves that AlCl4-.

One of the aluminium-chlorine bonds breaks and both electrons from it are used to join to the hydrogen. Removing the hydrogen from the ring forms the HCl which is also produced in the reaction, and the aluminium chloride catalyst is re-generated. The electrons which originally joined the hydrogen to the ring are now used to re-establish the delocalised system. THE SULPHONATION OF BENZENE

This page gives you the facts and a simple, uncluttered mechanism for the electrophilic substitution reaction between benzene and sulphuric acid (or sulphur trioxide). If you want this mechanism explained to you in detail, there is a link at the bottom of the page.

The electrophilic substitution reaction between benzene and sulphuric acid The facts There are two equivalent ways of sulphonating benzene:

Heat benzene under reflux with concentrated sulphuric acid for several hours. Warm benzene under reflux at 40C with fuming sulphuric acid for 20 to 30 minutes.

Or:

The product is benzenesulphonic acid. The electrophile is actually sulphur trioxide, SO3, and you may find the equation for the sulphonation reaction written:

Note: Which version of this equation you use will depend on what question you are being asked. If the question refers to the reaction with sulphuric acid, then you must use that one. If the question refers to SO3 as the electrophile, then you could use this one.

The formation of the electrophile The sulphur trioxide electrophile arises in one of two ways depending on which sort of acid you are using. Concentrated sulphuric acid contains traces of SO3 due to slight dissociation of the acid.

Fuming sulphuric acid, H2S2O7, can be thought of as a solution of SO3 in sulphuric acid - and so is a much richer source of the SO3. Sulphur trioxide is an electrophile because it is a highly polar molecule with a fair amount of positive charge on the sulphur atom. It is this which is attracted to the ring electrons. The electrophilic substitution mechanism Stage one

Stage two

The second stage of the reaction involves a transfer of the hydrogen from the ring to the negative oxygen.

EXPLAINING THE SULPHONATION OF BENZENE

This page guides you through the mechanism for the electrophilic substitution reaction between benzene and sulphuric acid (or sulphur trioxide)

The electrophilic substitution reaction between benzene and sulphuric acid The formation of the electrophile The electrophile is sulphur trioxide, and this arises in one of two ways depending on which sort of acid you are using. Concentrated sulphuric acid contains traces of SO3 due to slight dissociation of the acid.

Dissociation: This is a reversible splitting up of a compound. In this case, the sulphuric acid splits into water and SO 3, and at the same time these combine back together again to make sulphuric acid. The overall effect is that concentrated sulphuric acid contains small amounts of SO3.

Fuming sulphuric acid, H2S2O7, can be thought of as a solution of SO3 in sulphuric acid - and so is a much richer source of the SO3.
Note: You could think of the formula as being essentially H2SO4.SO3.

Although sulphur trioxide isn't ionic, it is highly polar. The three oxygens are more electronegative than the sulphur and so draw electrons towards themselves. That leaves the sulphur atom fairly positively charged. It is this + sulphur atom which attacks the benzene ring.
Note: If you aren't sure about electronegativity and polar bonds you might like to follow this link. Use the BACK button on your browser to return to this page.

The electrophilic substitution mechanism Stage one Two electrons from the delocalised system are used to form a new bond with the slightly positive sulphur atom. Because those two electrons aren't a part of the delocalised system any longer, the delocalisation is partly broken, and in the process the ring gains a positive charge. To make room for the new bond between the ring and the sulphur, two of the electrons joining the sulphur to one of the oxygens are forced right out on to the oxygen atom, giving it a negative charge.

The hydrogen shown on the ring is the one which was already attached to that top carbon atom. We need to show it there because it has to be removed in the second stage.
Note: If you aren't sure about the use of curly arrows follow this link before you go on. Use the BACK button on your browser to return to this page.

Stage two This second step is different from all the other benzene electrophilic substitution reactions you might have already looked at on this site. This time there isn't a separate negative ion to remove the hydrogen atom from the ring. Instead it is removed by a lone pair on the negative oxygen atom.

The lone pair forms a bond with the hydrogen atom, releasing the electrons in the hydrogen-to-ring bond so that they can reestablish the delocalisation. This particular mechanism needs to be drawn with rather more thought than the other electrophilic substitution mechanisms. In particular, you have to make sure that you put the negative charge on the right oxygen in the intermediate ion. It must be the

one which is closest to the hydrogen you intend to take off the ring, otherwise there is no way of drawing sensible curly arrows in the second stage!

ELECTROPHILIC SUBSTITUTION INTO ALREADY SUBSTITUTED BENZENE RINGS

This page discusses the problems which arise if you try to write the mechanism for an electrophilic substitution reaction into a benzene ring which already has something else attached to it. There are two problems you might come across:

Whereabouts in the ring does the substitution happen? Does this make a difference to how you draw the mechanisms? Can the group already attached to the ring get involved in any way?

Important! This page assumes that you can already write the mechanisms for substitution into a simple benzene ring. If you can't, go back to the electrophilic substitution menu and read about the reactions you are interested in before you tackle this page.

Electrophilic substitution in methylbenzene


The nitration of methylbenzene If you substitute a nitro group, -NO2, into the benzene ring in methylbenzene, you could possibly get any of the following products:

The carbon with the methyl group attached is thought of as the number 1 carbon, and the ring is then numbered around from 1 to 6. You number in a direction (in this case, clockwise) which produces the smaller number in the name - hence 2nitromethylbenzene rather than 6-nitromethylbenzene. In the case of methylbenzene, whatever you attach to the ring, you always get a mixture consisting mainly of the 2- and 4isomers. The methyl group is said to be 2,4-directing, in the sense that it seems to "push" incoming groups into those positions.
Isomers: Molecules which have the same molecular formula (i.e. contain exactly the same number and type of atoms), but with a different spatial arrangement of those atoms.

Some other groups which might already be on the ring (for example, the -NO2 group in nitrobenzene) "push" incoming groups into the 3- position. We'll have a quick look at this later on this page. What to expect in exams At A' level, you will not be expected to explain why different groups have different directing effects. It is possible, however, that you may be expected to remember the directing effect of one or two groups. It is very difficult to tell this from the syllabuses, most of which don't specifically exclude the possibility - but that could simply be an oversight. Your best option is to check past exam papers, or any support material published by your Exam Board. If in doubt, contact them and find out exactly what they expect.
Note: You will find links to help you to contact your Exam Board on the syllabuses page.

In an exam, then, you may have to remember the directing effect of a particular group, or more likely you will be told it. Your problem may then be to write the mechanism. How to write the mechanism for the nitration of methylbenzene
Important! This assumes that you are already familiar with the nitration of benzene. If you aren't, read about it before you go any further.

Reacting methylbenzene with a mixture of concentrated nitric and sulphuric acids gives both 2-nitromethylbenzene and 4nitromethylbenzene. The mechanism is exactly the same as the nitration of benzene. You just have to be careful about the way that you draw the structure of the intermediate ion. Making 2-nitromethylbenzene (the first step)

This just shows the first step of the electrophilic substitution reaction. Notice that the partial delocalisation in the intermediate ion covers all the carbon atoms in the ring except for the one that the -NO2 group gets attached to. That is the only point of interest in this example - everything else is just the same as with the nitration of benzene. The hydrogen atom is then removed by an HSO4- ion - exactly as in the benzene case. Making 4-nitromethylbenzene (the first step)

Once again, the only point of interest is in the way the partial delocalisation in the intermediate ion is drawn - again, it covers all the carbon atoms in the ring apart from the one with the -NO2 group attached. The electrophilic substitution reaction between methylbenzene and chlorine This is a good example of a case where what is already attached to the ring can also get involved in the reaction. It is possible to get two quite different substitution reactions between methylbenzene and chlorine depending on the conditions used. The chlorine can substitute into the ring or into the methyl group. Here we are only interested in substitution into the ring. This happens in the presence of aluminium chloride or iron, and in the absence of UV light.
Note: If you are also interested in substitution into the methyl group (in the presence of UV light - and with no catalyst present), you will find this explained in the page on the free radical reaction between methylbenzene and chlorine.

Substituting into the ring gives a mixture of 2chloromethylbenzene and 4-chloromethylbenzene.

The mechanisms are exactly the same as the substitution of chlorine into benzene - although you would have to be careful about the way you draw the intermediate ion.
Important! This assumes that you are already familiar with the chlorination of benzene. If you aren't, read about it before you go any further.

For example, the complete mechanism for substitution into the 4- position is: Stage one

Stage two

Electrophilic substitution in nitrobenzene


Substitution into the 3- position (the first step) Methyl groups direct new groups into the 2- and 4- positions, but

a nitro group, -NO2, already on the ring directs incoming groups into the 3- position. For example, if the temperature is raised above 50C, the nitation of benzene doesn't just produce nitrobenzene - it also produces some 1,3-dinitrobenzene. A second nitro group is substituted into the ring in the 3- position.

The mechanism is exactly the same as the nitration of benzene or of methylbenzene - you just have to be careful in drawing the intermediate ion. Draw the partial delocalisation to include all the carbons except for the one the new -NO2 group gets attached to.

In the second stage, the hydrogen atom is then removed by an HSO4- ion - exactly as in the benzene case. This isn't shown because there's nothing new.

NUCLEOPHILIC SUBSTITUTION MECHANISMS MENU


Types of halogenoalkanes . . . Desribes what is meant by primary, secondary and tertiary halogenoalkanes. Essential knowledge if you are to make sense of everything else in this section. What is nucleophilic substitution? . . .

Includes background material on the bonding in halogenoalkanes, and general mechanisms for their nucleophilic substitution reactions. Substitution reactions involving hydroxide ions . . . The mechanisms for the formation of alcohols from halogenoalkanes by reaction with hydroxide ions. Substitution reactions involving water . . . The mechanisms for the formation of alcohols from halogenoalkanes by reaction with water. Substitution reactions involving cyanide ions . . . The mechanisms for the formation of nitriles from halogenoalkanes by reaction with cyanide ions. Substitution reactions involving ammonia . . . The mechanisms for the formation of various sorts of amines from halogenoalkanes by reaction with ammonia. Also includes the reactions between halogenoalkanes and amines.

HALOGENOALKANES
Halogenoalkanes are also called haloalkanes or alkyl halides. All halogenoalkanes contain a halogen atom - fluorine, chlorine, bromine or iodine - attached to an alkyl group.
Note: An alkyl group is a group such as methyl, CH 3, or ethyl, CH3CH2. These are groups containing chains of carbon atoms which may be branched. Alkyl groups are given the general symbol R.

The different kinds of halogenoalkanes


Primary halogenoalkanes In a primary (1) halogenoalkane, the carbon which carries the

halogen atom is only attached to one other alkyl group. Some examples of primary halogenoalkanes include:

Notice that it doesn't matter how complicated the attached alkyl group is. In each case there is only one linkage from the CH2 group holding the halogen to an alkyl group. There is an exception to this. CH3Br and the other methyl halides are often counted as primary halogenoalkanes even though there are no alkyl groups attached to the carbon with the halogen on it. Secondary halogenoalkanes In a secondary (2) halogenoalkane, the carbon with the halogen attached is joined directly to two other alkyl groups, which may be the same or different. Examples:

Tertiary halogenoalkanes In a tertiary (3) halogenoalkane, the carbon atom holding the halogen is attached directly to three alkyl groups, which may be any combination of same or different. Examples:

NUCLEOPHILIC SUBSTITUTION Background Bonding in the halogenoalkanes Halogenoalkanes (also known as haloalkanes or

alkyl halides) are compounds containing a halogen atom (fluorine, chlorine, bromine or iodine) joined to one or more carbon atoms in a chain. The interesting thing about these compounds is the carbon-halogen bond, and all the nucleophilic substitution reactions of the halogenoalkanes involve breaking that bond. The polarity of the carbon-halogen bonds With the exception of iodine, all of the halogens are more electronegative than carbon. Electronegativity values (Pauling scale)
C 2.5 F Cl Br I 4.0 3.0 2.8 2.5

Note: If you aren't sure about electronegativity and bond polarity follow this link before you read on. Use the BACK button on your browser to return to this page.

That means that the electron pair in the carbon-halogen bond will be dragged towards the halogen end, leaving the halogen slightly negative ( -) and the carbon slightly positive ( +) except in the carbon-iodine case. Although the carbon-iodine bond doesn't have a permanent dipole, the bond is very easily polarised by anything approaching it. Imagine a negative ion approaching the bond from the far side of the carbon atom:

The fairly small polarity of the carbon-bromine bond will be

increased by the same effect. The strengths of the carbon-halogen bonds


Note: If you haven't done any work on bond strengths, or are a bit rusty, it doesn't matter. Just realise that the bigger the number, the stronger the bond. And don't worry if you have found slightly different numbers in a different data source - there is a lot of variability in the quoted values, but the overall pattern is still the same.

Look at the strengths of various bonds (all values in kJ mol-1). C-H 413 C-F C-Cl C-Br C-I 467 346 290 228

In all of these nucleophilic substitution reactions, the carbonhalogen bond has to be broken at some point during the reaction. The harder it is to break, the slower the reaction will be. The carbon-fluorine bond is very strong (stronger than C-H) and isn't easily broken. It doesn't matter that the carbon-fluorine bond has the greatest polarity - the strength of the bond is much more important in determining its reactivity. You might therefore expect fluoroalkanes to be very unreactive - and they are! We shall simply ignore them from now on. In the other halogenoalkanes, the bonds get weaker as you go from chlorine to bromine to iodine. That means that chloroalkanes react most slowly, bromoalkanes react faster, and iodoalkanes react faster still. Rates of reaction: RCl < RBr < RI Where "<" is read as "is less than" - or, in this instance, "is slower than", and R represents any alkyl group.

Warning! Before you read on it is essential that you know

exactly what your syllabus says about these reactions so that you can extract the right amount of detail from what follows. The problem is that there are two different mechanisms depending on the type of halogenoalkane you are using (whether primary, secondary or tertiary). Some syllabuses try to make things simpler by restricting you to just one of the two mechanisms. Where the syllabus is vague, look at recent exam papers and mark schemes, or any support material published by your examiners. Haven't got a syllabus or recent exam papers? If you are working to one of the UK-based syllabuses for 16 - 18 year olds, follow this link to find out how to get them. You must know what your examiners expect in this topic.

Nucleophilic substitution in primary halogenoalkanes


You will need to know about this if your syllabus talks about "primary halogenoalkanes" or about SN2 reactions. If the syllabus is vague, check recent exam papers and mark schemes, and compare them against what follows. Nucleophiles A nucleophile is a species (an ion or a molecule) which is strongly attracted to a region of positive charge in something else. Nucleophiles are either fully negative ions, or else have a strongly - charge somewhere on a molecule. Common nucleophiles are hydroxide ions, cyanide ions, water and ammonia.

Notice that each of these contains at least one lone pair of electrons, either on an atom carrying a full negative charge, or on a very electronegative atom carrying a substantial - charge. The nucleophilic substitution reaction - an SN2 reaction

We'll talk this mechanism through using an ion as a nucleophile, because it's slightly easier. The water and ammonia mechanisms involve an extra step which you can read about on the pages describing those particular mechanisms. We'll take bromoethane as a typical primary halogenoalkane. The bromoethane has a polar bond between the carbon and the bromine. We'll look at its reaction with a general purpose nucleophilic ion which we'll call Nu-. This will have at least one lone pair of electrons. Nu- could, for example, be OH- or CN-.

The lone pair on the Nu- ion will be strongly attracted to the + carbon, and will move towards it, beginning to make a coordinate (dative covalent) bond. In the process the electrons in the C-Br bond will be pushed even closer towards the bromine, making it increasingly negative.
Note: A co-ordinate bond is a covalent bond in which both electrons come from one of the atoms.

The movement goes on until the -Nu is firmly attached to the carbon, and the bromine has been expelled as a Br- ion.
Note: We haven't shown all the lone pairs on the bromine. These other lone pairs aren't involved in the reaction, and including them simply clutters the diagram to no purpose.

Things to notice The Nu- ion approaches the + carbon from the side away from the bromine atom. The large bromine atom hinders attack from its side and, being -, would repel the incoming Nu- anyway. This attack from the back is important if you need to understand

why tertiary halogenoalkanes have a different mechanism. We'll discuss this later on this page. There is obviously a point in which the Nu- is half attached to the carbon, and the C-Br bond is half way to being broken. This is called a transition state. It isn't an intermediate. You can't isolate it - even for a very short time. It's just the mid-point of a smooth attack by one group and the departure of another. How to write the mechanism The simplest way is:

Note: In exam you must show the lone pair of electrons on the nucleophile (in this case, the Nu ion). It probably doesn't matter whether you show them on the departing Br ion or not. If you aren't happy about the use of curly arrows in mechanisms, follow this link before you go on. Use the BACK button on your browser to return to this page.

Technically, this is known as an SN2 reaction. S stands for substitution, N for nucleophilic, and the 2 is because the initial stage of the reaction involves two species - the bromoethane and the Nu- ion. If your syllabus doesn't refer to SN2 reactions by name, you can just call it nucleophilic substitution.

Some examiners like you to show the transition state in the mechanism, in which case you need to write it in a bit more detail - showing how everything is arranged in space.

Be very careful when you draw the transition state to make a clear difference between the dotted lines showing the half-made and half-broken bonds, and those showing the bonds going back into the paper. Notice that the molecule has been inverted during the reaction rather like an umbrella being blown inside-out.
Note: If you aren't happy about the various ways of drawing bonds, it is important to follow this link to find out exactly what the various symbols mean. It is also important to know which of these ways of drawing the mechanism your particular examiners want you to use. If you haven't already checked your syllabus, recent exam papers and mark schemes, you must do so! At the time of writing, Edexcel, for example, wanted the transition state included, and that isn't obvious from their syllabus. You have to check mark schemes and examiners reports. Use the BACK button on your browser to return to this page.

Nucleophilic substitution in tertiary halogenoalkanes


Warning! Check your syllabus, past papers and any support material published by your examiners to find out whether you need this. If there's no mention of tertiary halogenoalkanes or SN1 reactions, then you probably don't

need it.

Remember that a tertiary halogenoalkane has three alkyl groups attached to the carbon with the halogen on it. These alkyl groups can be the same or different, but in this section, we shall just consider a simple one, (CH3)3CBr - 2-bromo-2-methylpropane. The nucleophilic substitution reaction - an SN1 reaction Once again, we'll talk this mechanism through using an ion as a nucleophile, because it's slightly easier, and again we'll look at the reaction of a general purpose nucleophilic ion which we'll call Nu-. This will have at least one lone pair of electrons. Why is a different mechanism necessary? You will remember that when a nucleophile attacks a primary halogenoalkane, it approaches the + carbon atom from the side away from the halogen atom. With a tertiary halogenoalkane, this is impossible. The back of the molecule is completely cluttered with CH3 groups.

Since any other approach is prevented by the bromine atom, the reaction has to go by an alternative mechanism. The alternative mechanism</I.< b>
Important! To understand this section, you need to know what a carbocation (carbonium ion) is, and about the relative stabilities of primary, secondary and tertiary carbocations. If you follow this link, use the BACK button on your browser to return to this page.

The reaction happens in two stages. In the first, a small proportion of the halogenoalkane ionises to give a carbocation and a bromide ion.

This reaction is possible because tertiary carbocations are relatively stable compared with secondary or primary ones. Even so, the reaction is slow. Once the carbocation is formed, however, it would react immediately it came into contact with a nucleophile like Nu-. The lone pair on the nucleophile is strongly attracted towards the positive carbon, and moves towards it to create a new bond.

How fast the reaction happens is going to be governed by how fast the halogenoalkane ionises. Because this initial slow step only involves one species, the mechanism is described as S N1 substitution, nucleophilic, one species taking part in the initial slow step. Why don't primary halogenoalkanes use the SN1 mechanism? If a primary halogenoalkane did use this mechanism, the first step would be, for example:

A primary carbocation would be formed, and this is much more energetically unstable than the tertiary one formed from tertiary halogenoalkanes - and therefore much more difficult to produce. This instability means that there will be a very high activation energy for the reaction involving a primary halogenoalkane. The

activation energy is much less if it undergoes an SN2 reaction and so that's what it does instead.

Nucleophilic substitution in secondary halogenoalkanes


There isn't anything new in this. Secondary halogenoalkanes will use both mechanisms some molecules will react using the SN2 mechanism and others the SN1.

The SN2 mechanism is possible because the back of the molecule isn't completely cluttered by alkyl groups and so the approaching nucleophile can still get at the + carbon atom. The SN1 mechanism is possible because the secondary carbocation formed in the slow step is more stable than a primary one. It isn't as stable as a tertiary one though, and so the SN1 route isn't as effective as it is with tertiary halogenoalkanes.

THE NUCLEOPHILIC SUBSTITUTION REACTIONS BETWEEN HALOGENOALKANES AND HYDROXIDE IONS


This page gives you the facts and simple, uncluttered mechanisms for the nucleophilic substitution reactions between halogenoalkanes and hydroxide ions (from, for example, sodium hydroxide). If you want the mechanisms explained to you in detail, there is a link at the bottom of the page.

The reaction of primary halogenoalkanes with hydroxide ions

Important! If you aren't sure about the difference between primary, secondary and tertiary halogenoalkanes, it is essential that you follow this link before you go on. Use the BACK button on your browser to return to this page.

The facts If a halogenoalkane is heated under reflux with a solution of sodium or potassium hydroxide, the halogen is replaced by -OH and an alcohol is produced. Heating under reflux means heating with a condenser placed vertically in the flask to prevent loss of volatile substances from the mixture. The solvent is usually a 50/50 mixture of ethanol and water, because everything will dissolve in that. The halogenoalkane is insoluble in water. If you used water alone as the solvent, the halogenoalkane and the sodium hydroxide solution wouldn't mix and the reaction could only happen where the two layers met. For example, using 1-bromopropane as a typical primary halogenoalkane:

You could write the full equation rather than the ionic one, but it slightly obscures what's going on:

The bromine (or other halogen) in the halogenoalkane is simply replaced by an -OH group - hence a substitution reaction. In this example, propan-1-ol is formed. The mechanism Here is the mechanism for the reaction involving bromoethane:

This is an example ofnucleophilic substitution. Because the mechanism involves collision between two species in the slow step (in this case, the only step) of the reaction, it is known as an SN2 reaction.
Note: Unless your syllabus specifically mentions SN2 by name, you can just call it nucleophilic substitution.

If your examiners want you to show the transition state, draw the mechanism like this:

The reaction of tertiary halogenoalkanes with hydroxide ions


The facts The facts of the reaction are exactly the same as with primary halogenoalkanes. If the halogenoalkane is heated under reflux with a solution of sodium or potassium hydroxide in a mixture of ethanol and water, the halogen is replaced by -OH, and an alcohol is produced.

For example:

Or if you want the full equation rather than the ionic one:

The mechanism This mechanism involves an initial ionisation of the halogenoalkane:

followed by a very rapid attack by the hydroxide ion on the carbocation (carbonium ion) formed:

This is again an example ofnucleophilic substitution. This time the slow step of the reaction only involves one species - the halogenoalkane. It is known as an SN1 reaction.

The reaction of secondary halogenoalkanes with hydroxide ions


The facts The facts of the reaction are exactly the same as with primary or tertiary halogenoalkanes. The halogenoalkane is heated under reflux with a solution of sodium or potassium hydroxide in a mixture of ethanol and water. For example:

The mechanism Secondary halogenoalkanes use both SN2 and SN1 mechanisms. For example, the SN2 mechanism is:

Should you need it, the two stages of the SN1 mechanism are:

Note: There is another reaction between halogenoalkanes and hydroxide ions involving an elimination reaction. If you want to explore that as an alternative, follow this link.

EXPLAINING THE NUCLEOPHILIC SUBSTITUTION REACTIONS BETWEEN HALOGENOALKANES AND HYDROXIDE IONS
This page guides you through the nucleophilic substitution mechanisms for the reactions between halogenoalkanes and hydroxide ions from, for example, sodium hydroxide.
Important! It would help if you first read the page What is

nucleophilic substitution? before you go on. You must also be clear about the differences between primary, secondary and tertiary halogenoalkanes.

The reactions between primary or secondary halogenoalkanes and hydroxide ions - the SN2 mechanism
Hydroxide ions as nucleophiles A nucleophile is a species (an ion or a molecule) which is strongly attracted to a region of positive charge in something else. Nucleophiles are either fully negative ions, or else have a strongly - charge somewhere on a molecule. In the case of the hydroxide ion, there is a full negative charge on the oxygen, as well as three lone pairs of electrons. The nucleophilic substitution reaction - an SN2 reaction We'll talk this reaction through with a primary halogenoalkane to start with, taking bromoethane as typical. The bromoethane has a polar bond between the carbon and the bromine.

Note: In an exam you must show the lone pair of electrons on the nucleophile (in this case, the OH ion). It probably doesn't matter whether you show them on the departing Br ion or not.

One of the lone pairs on the OH- ion will be strongly attracted to the + carbon, and will move towards it, beginning to form a

bond with it. The approaching negative ion will repel the electrons in the carbon-bromine bond closer and closer to the bromine. At some point during this, the -OH group and the bromine will both be half-attached to the carbon. This is called the transition state for the reaction. It isn't an intermediate - you can't isolate it and it doesn't have any independent existence. It's just the halfway stage of a smooth movement of atoms and electrons. The movement goes on until the -OH is firmly attached to the carbon, and the bromine has been expelled as a Br- ion. You may need to show the formation of the intermediate in the mechanism (depending on what your examiners want). It simply needs you to draw the mechanism showing some more detail about how the various groups are arranged in space.

Be very careful when you draw the transition state to make a clear difference between the dotted lines showing the half-made and half-broken bonds, and those showing the bonds going back into the paper. Notice that the molecule has been inverted during the reaction rather like an umbrella being blown inside-out.
Note: If you aren't happy about the various ways of drawing bonds, it is important to follow this link to find out exactly what the various symbols mean. It is also important to know which of these ways of drawing the

mechanism your particular examiners want you to use. If you haven't already checked your syllabus, recent exam papers and mark schemes, you must do so! At the time of writing, Edexcel, for example, wanted the transition state included, and that isn't obvious from their syllabus. You have to check mark schemes and examiners reports. Use the BACK button on your browser to return to this page.

Technically, this is known as an SN2 reaction. S stands for substitution, N for nucleophilic, and the 2 is because the initial stage of the reaction involves two species - the bromoethane and the OH- ion. If your syllabus doesn't refer to SN2 reactions by name, you can just call it nucleophilic substitution. The SN2 reaction in secondary halogenoalkanes The reaction can happen in exactly the same way with a secondary halogenoalkane, although they also have the potential for reacting via a different mechanism (which we'll deal with shortly).

Again, a lone pair on the approaching hydroxide ion forms a bond with the + carbon and, in the process, the electrons in the carbon-bromine bond are forced entirely onto the bromine to create a bromide ion.

The reactions between secondary or tertiary halogenoalkanes and hydroxide ions - the SN1 mechanism
Note: Are you sure your syllabus wants this? Several syllabuses restrict you to primary halogenoalkanes. Haven't got a syllabus? If you are working to a UK-based syllabus for 16 - 18 year olds, follow this link to find out how to

get one.

To start with, we'll talk this mechanism through with a simple tertiary halogenoalkane like the one on the right (2-bromo-2-methylpropane).

Why do tertiary halogenoalkanes need a different mechanism? When a nucleophile attacks a primary halogenoalkane, it approaches the + carbon atom from the side away from the halogen atom. Any other approach is prevented by the halogen atom, which is both bulky and slightly negatively charged. The charge repels the incoming nucleophile. With a tertiary halogenoalkane, this approach from the back is impossible. The back of the molecule is completely cluttered with CH3 groups. The SN1 mechanism The reaction happens in two stages. In the first, a small proportion of the halogenoalkane ionises to give a carbocation (carbonium ion) and a bromide ion.

This reaction is possible because tertiary carbocations are relatively stable compared with secondary or primary ones. Even so, the reaction is slow.
Note: Not sure about the stability of carbocations (carbonium ions)? Follow this link if you need to find out.

Once the carbocation is formed, however, it would react immediately it came into contact with an OH- ion. The lone pair on the nucleophile is strongly attracted towards the positive

carbon, and moves towards it to create a new bond.

How fast the reaction happens is going to be governed by how fast the halogenoalkane ionises - because that's a slow process. Because this initial slow step only involves one species, the mechanism is described as SN1 - substitution, nucleophilic, one species taking part in the initial slow step. The SN1 mechanism in secondary halogenoalkanes Secondary halogenoalkanes (like 2-bromopropane) can use either the SN1 or the SN2 mechanism. The back of the molecule is rather more cluttered than in a primary halogenoalkane, but there is still room for the lone pair on the nucleophile to approach and form a bond. We've already dealt with that reaction. It is also possible to get some slight ionisation of the halogenoalkane to give an SN1 mechanism, but this reaction is much less successful than with tertiary halogenoalkanes, because the secondary carbocation formed isn't as stable as a tertiary one.

Once the carbocation has been formed, it will react immediately with a hydroxide ion. A lone pair on the hydroxide ion is strongly attracted to the positive carbon, moves towards it, and forms a bond.

THE NUCLEOPHILIC SUBSTITUTION REACTIONS BETWEEN HALOGENOALKANES AND WATER

This page gives you the facts and simple, uncluttered mechanisms for the nucleophilic substitution reactions between halogenoalkanes and water. If you want the mechanisms explained to you in detail, there is a link at the bottom of the page.

The reaction of primary halogenoalkanes with water


Important! If you aren't sure about the difference between primary, secondary and tertiary halogenoalkanes, it is essential that you follow this link before you go on. Use the BACK button on your browser to return to this page.

The facts There is only a slow reaction between a primary halogenoalkane and water even if they are heated. The halogen atom is replaced by -OH. For example, using 1-bromoethane as a typical primary halogenoalkane:

An alcohol is produced together with hydrobromic acid. Be careful not to call this hydrogen bromide. Hydrogen bromide is a gas. When it is dissolved it in water (as it will be here), it's called hydrobromic acid. The mechanism The mechanism involves two steps. The first is a simple nucleophilic substitution reaction:

Because the mechanism involves collision between two species in this slow step of the reaction, it is known as an SN2 reaction.
Note: Unless your syllabus specifically mentions SN2 by name, you can just call it nucleophilic substitution.

The nucleophilic substitution is very slow because water isn't a very good nucleophile. It lacks the full negative charge of, say, a hydroxide ion. The second step of the reaction simply tidies up the product. A water molecule removes one of the hydrogens attached to the oxygen to give an alcohol and a hydroxonium ion (also known as a hydronium ion or an oxonium ion).

The hydroxonium ion and the bromide ion (from the nucleophilic substitution stage of the reaction) make up the hydrobromic acid which is formed as well as the alcohol.

The reaction of tertiary halogenoalkanes with water


The facts If the halogenoalkane is heated under reflux with water, the halogen is replaced by -OH to give an alcohol. Heating under reflux means heating with a condenser placed vertically in the flask to prevent loss of volatile substances from the mixture. The reaction happens much faster than the corresponding one involving a primary halogenoalkane. For example:

The mechanism This mechanism involves an initial ionisation of the halogenoalkane:

followed by a very rapid attack by the water on the carbocation (carbonium ion) formed:

This is again an example ofnucleophilic substitution. This time the slow step of the reaction only involves one species - the halogenoalkane. It is known as an SN1 reaction. Now there is a final stage in which the product is tidied up. A water molecule removes one of the hydrogens attached to the oxygen to give an alcohol and a hydroxonium ion - exactly as happens with primary halogenoalkanes.

The rate of the overall reaction is governed entirely by how fast the halogenoalkane ionises. The fact that water isn't as good a nucleophile as, say, OH- doesn't make any difference. The water isn't involved in the slow step of the reaction.

The reaction of secondary halogenoalkanes with

water
It is very unlikely that any of the current UK-based syllabuses for 16 - 18 year olds will ask you about this. In the extremely unlikely event that you will ever need it, secondary halogenoalkanes use both an SN2 mechanism and an SN1. Make sure you understand what happens with primary and tertiary halogenoalkanes, and then adapt it for secondary ones should ever need to.

EXPLAINING THE NUCLEOPHILIC SUBSTITUTION REACTIONS BETWEEN HALOGENOALKANES AND WATER

This page guides you through the nucleophilic substitution mechanisms for the reactions between halogenoalkanes and water. It deals only with primary and tertiary halogenoalkanes. No current UK-based syllabus for 16 - 18 year olds is likely to ask about the reaction between secondary halogenoalkanes and water. (It's not difficult - it's just not there!)
Important! It would help if you first read the page What is nucleophilic substitution? before you go on. You must also be clear about the differences between primary, secondary and tertiary halogenoalkanes.

The reaction between primary halogenoalkanes and water - the SN2 mechanism
Water as a nucleophile A nucleophile is a species (an ion or a molecule) which is strongly attracted to a region of positive charge in something else. Nucleophiles are either fully negative ions, or else have a strongly - charge somewhere on a molecule.

Water obviously doesn't carry a negative charge. However, oxygen is much more electronegative than hydrogen, and so the oxygen atom has a fairly substantial - charge to back up its two lone pairs.
Note: If you aren't sure about electronegativity follow this link before you read on. Use the BACK button on your browser to return to this page.

The attack on the halogenoalkane is therefore by one of the lone pairs on the oxygen. Because there isn't a full negative charge, water isn't going to be as good a nucleophile as a negative ion like OH-, and so the reaction is slower. The nucleophilic substitution reaction - an SN2 reaction We'll talk this reaction through with bromoethane as a typical primary halogenoalkane. The bromoethane has a polar bond between the carbon and the bromine.

One of the lone pairs on the water will be strongly attracted to the + carbon, and will move towards it, beginning to form a bond with it. The approaching lone pair will repel the electrons in the carbon-bromine bond closer and closer to the bromine. The movement goes on until the water is firmly attached to the carbon, and the bromine has been expelled as a Br- ion. Notice that the oxygen in the product ion carries a positive charge (highlighted in red to draw attention to it). That charge has to be there for two reasons:

On the left hand side of the equation you start with two

overall neutral molecules. Assuming you forgot about the positive charge, you would end up with a neutral species and a negative ion on the right. Charges must balance in equations, so something is wrong. The oxygen looks wrong! The oxygen atom is joined to 3 things rather than its usual 2. Oxygen can only join to 3 things if it carries a positive charge.

Note: Oxygen with a positive charge has the same arrangement of electrons as a nitrogen atom - which normally forms 3 bonds.

Technically, this is known as an SN2 reaction. S stands for substitution, N for nucleophilic, and the 2 is because the initial stage of the reaction involves two species - the bromoethane and the water. If your syllabus doesn't refer to SN2 reactions by name, you can just call it nucleophilic substitution. Finally a hydrogen ion is pulled off the product ion by another water molecule from the solution. A lone pair on the new water molecule forms a bond with a hydrogen atom, forcing the bonding pair of electrons back on to the positive oxygen. That cancels the positive charge.

The organic product is ethanol. The formula is distorted in this diagram so that you can clearly see the relationship between the atoms on both sides of the equation. The other product is a hydroxonium ion (also known as a hydronium ion or an oxonium ion). This is just a hydrogen ion attached to a water molecule - what is often written as H+(aq).

The reaction between tertiary halogenoalkanes and water - the SN1 mechanism
We'll talk this mechanism through with a simple tertiary halogenoalkane like the one on the right

(2-bromo-2-methylpropane).

Why do tertiary halogenoalkanes need a different mechanism? When a nucleophile attacks a primary halogenoalkane, it approaches the + carbon atom from the side away from the halogen atom. Any other approach is prevented by the halogen atom, which is both bulky and slightly negatively charged. The charge repels the incoming nucleophile. With a tertiary halogenoalkane, this approach from the back is impossible. The back of the molecule is completely cluttered with CH3 groups. The SN1 mechanism In the first stage, a small proportion of the halogenoalkane ionises to give a carbocation (carbonium ion) and a bromide ion.

This reaction is possible because tertiary carbocations are relatively stable compared with secondary or primary ones. Even so, the reaction is slow.
Note: Not sure about the stability of carbocations (carbonium ions)? Follow this link if you need to find out.

Once the carbocation is formed, however, it would react immediately it came into contact with a water molecule. One of the lone pairs on the water is strongly attracted towards the positive carbon, and moves towards it to create a new bond.

How fast the reaction happens overall is going to be governed

by how fast the halogenoalkane ionises - because that's a slow process. Because this initial slow step only involves one species, the mechanism is described as SN1 - substitution, nucleophilic, one species taking part in the initial slow step. The water takes part in the fast step of the reaction, and so the fact that a weakish nucleophile like water is involved doesn't significantly slow the overall reaction down. The rate is determined by the slow ionisation of the halogenoalkane. As with primary halogenalkanes, there is a final stage to this reaction in which a hydrogen ion is transferred from the organic ion to a water molecule in the solution. What happens is exactly the same as with the primary halogenoalkanes described above.

THE NUCLEOPHILIC SUBSTITUTION REACTIONS BETWEEN HALOGENOALKANES AND CYANIDE IONS

This page gives you the facts and simple, uncluttered mechanisms for the nucleophilic substitution reactions between halogenoalkanes and cyanide ions (from, for example, potassium cyanide). If you want the mechanisms explained to you in detail, there is a link at the bottom of the page.

The reaction of primary halogenoalkanes with cyanide ions


Important! If you aren't sure about the difference between primary, secondary and tertiary halogenoalkanes, it is essential that you follow this link before you go on.

Use the BACK button on your browser to return to this page.

The facts If a halogenoalkane is heated under reflux with a solution of sodium or potassium cyanide in ethanol, the halogen is replaced by a -CN group and a nitrile is produced. Heating under reflux means heating with a condenser placed vertically in the flask to prevent loss of volatile substances from the mixture. The solvent is important. If water is present you tend to get substitution by -OH instead of -CN.
Note: A solution of potassium cyanide in water is quite alkaline, and contains significant amounts of hydroxide ions. These react with the halogenoalkane.

For example, using 1-bromopropane as a typical primary halogenoalkane:

You could write the full equation rather than the ionic one, but it slightly obscures what's going on:

The bromine (or other halogen) in the halogenoalkane is simply replaced by a -CN group - hence a substitution reaction. In this example, butanenitrile is formed.
Note: When you are naming nitriles, you have to remember to include the carbon in the -CN group when you count the longest chain. In this example, there are 4 carbons in the longest chain - hence butanenitrile.

The mechanism

Here is the mechanism for the reaction involving bromoethane:

This is an example ofnucleophilic substitution. Because the mechanism involves collision between two species in the slow step (in this case, the only step) of the reaction, it is known as an SN2 reaction.
Note: Unless your syllabus specifically mentions SN2 by name, you can just call it nucleophilic substitution.

If your examiners want you to show the transition state, draw the mechanism like this:

The reaction of tertiary halogenoalkanes with cyanide ions


The facts The facts of the reaction are exactly the same as with primary halogenoalkanes. If the halogenoalkane is heated under reflux with a solution of sodium or potassium cyanide in ethanol, the

halogen is replaced by -CN, and a nitrile is produced. For example:

Or if you want the full equation rather than the ionic one:

The mechanism This mechanism involves an initial ionisation of the halogenoalkane:

followed by a very rapid attack by the cyanide ion on the carbocation (carbonium ion) formed:

This is again an example ofnucleophilic substitution. This time the slow step of the reaction only involves one species - the halogenoalkane. It is known as an SN1 reaction.

The reaction of secondary halogenoalkanes with cyanide ions


The facts The facts of the reaction are exactly the same as with primary or tertiary halogenoalkanes. The halogenoalkane is heated under reflux with a solution of sodium or potassium cyanide in ethanol. For example:

The mechanism Secondary halogenoalkanes use both SN2 and SN1 mechanisms. For example, the SN2 mechanism is:

Should you need it, the two stages of the SN1 mechanism are:

EXPLAINING THE NUCLEOPHILIC SUBSTITUTION REACTIONS BETWEEN HALOGENOALKANES AND CYANIDE IONS

This page guides you through the nucleophilic substitution mechanisms for the reactions between halogenoalkanes and cyanide ions from, for example, potassium cyanide.
Important! It would help if you first read the page What is nucleophilic substitution? before you go on. You must also be clear about the differences between primary, secondary and tertiary halogenoalkanes.

The reactions between primary or secondary halogenoalkanes and cyanide ions - the SN2

mechanism
Cyanide ions as nucleophiles A nucleophile is a species (an ion or a molecule) which is strongly attracted to a region of positive charge in something else. Nucleophiles are either fully negative ions, or else have a strongly - charge somewhere on a molecule. In the case of the cyanide ion, there is a full negative charge on the carbon, as well as a lone pair of electrons.
Note: There is a lone pair on the nitrogen atom as well, but this isn't shown to avoid confusion. The one on the carbon is more important in this instance because that's where the negative charge is. The combination of the lone pair and the negative charge makes the carbon end of the ion the nucleophile.

The nucleophilic substitution reaction - an SN2 reaction We'll talk this reaction through with a primary halogenoalkane to start with, taking bromoethane as typical. The bromoethane has a polar bond between the carbon and the bromine.

Note: In an exam you must show the lone pair of electrons on the nucleophile (in this case, the CN ion). It probably doesn't matter whether you show them on the departing Br ion or not.

The lone pair on the carbon of the cyanide ion will be strongly attracted to the + carbon, and will move towards it, beginning to form a bond with it. The approaching negative ion will repel the electrons in the carbon-bromine bond closer and closer to the bromine.

At some point during this, the -CN group and the bromine will both be half-attached to the carbon. This is called the transition state for the reaction. It isn't an intermediate - you can't isolate it and it doesn't have any independent existence. It's just the halfway stage of a smooth movement of atoms and electrons. The movement goes on until the -CN is firmly attached to the carbon, and the bromine has been expelled as a Br- ion. You may need to show the formation of the intermediate in the mechanism (depending on what your examiners want). It simply needs you to draw the mechanism showing some more detail about how the various groups are arranged in space.

Be very careful when you draw the transition state to make a clear difference between the dotted lines showing the half-made and half-broken bonds, and those showing the bonds going back into the paper. Notice that the molecule has been inverted during the reaction rather like an umbrella being blown inside-out.
Note: If you aren't happy about the various ways of drawing bonds, it is important to follow this link to find out exactly what the various symbols mean. It is also important to know which of these ways of drawing the mechanism your particular examiners want you to use. If you haven't already checked your syllabus, recent exam papers and mark schemes, you must do so! At the time of writing, Edexcel, for example, wanted the transition state included, and that isn't obvious from their syllabus. You have to check mark schemes and

examiners reports. Use the BACK button on your browser to return to this page.

Technically, this is known as an SN2 reaction. S stands for substitution, N for nucleophilic, and the 2 is because the initial stage of the reaction involves two species - the bromoethane and the CN- ion. If your syllabus doesn't refer to SN2 reactions by name, you can just call it nucleophilic substitution. The SN2 reaction in secondary halogenoalkanes The reaction can happen in exactly the same way with a secondary halogenoalkane, although they also have the potential for reacting via a different mechanism (which we'll deal with shortly).

Again, the lone pair on the approaching cyanide ion forms a bond with the + carbon and, in the process, the electrons in the carbon-bromine bond are forced entirely onto the bromine to create a bromide ion.

The reactions between secondary or tertiary halogenoalkanes and cyanide ions - the SN1 mechanism
Note: Are you sure your syllabus wants this? Several syllabuses restrict you to primary halogenoalkanes. Haven't got a syllabus? If you are working to a UK-based syllabus for 16 - 18 year olds, follow this link to find out how to get one.

To start with, we'll talk this mechanism through with a simple tertiary halogenoalkane like the one on the right (2-bromo-2-methylpropane).

Why do tertiary halogenoalkanes need a different mechanism? When a nucleophile attacks a primary halogenoalkane, it approaches the + carbon atom from the side away from the halogen atom. Any other approach is prevented by the halogen atom, which is both bulky and slightly negatively charged. The charge repels the incoming nucleophile. With a tertiary halogenoalkane, this approach from the back is impossible. The back of the molecule is completely cluttered with CH3 groups. The SN1 mechanism The reaction happens in two stages. In the first, a small proportion of the halogenoalkane ionises to give a carbocation (carbonium ion) and a bromide ion.

This reaction is possible because tertiary carbocations are relatively stable compared with secondary or primary ones. Even so, the reaction is slow.
Note: Not sure about the stability of carbocations (carbonium ions)? Follow this link if you need to find out.

Once the carbocation is formed, however, it would react immediately it came into contact with a CN- ion. The lone pair on the nucleophile is strongly attracted towards the positive carbon, and moves towards it to create a new bond.

How fast the reaction happens is going to be governed by how fast the halogenoalkane ionises - because that's a slow process. Because this initial slow step only involves one species, the mechanism is described as SN1 - substitution, nucleophilic, one species taking part in the initial slow step. The SN1 mechanism in secondary halogenoalkanes Secondary halogenoalkanes (like 2-bromopropane) can use either the SN1 or the SN2 mechanism. The back of the molecule is rather more cluttered than in a primary halogenoalkane, but there is still room for the lone pair on the nucleophile to approach and form a bond. We've already dealt with that reaction. It is also possible to get some slight ionisation of the halogenoalkane to give an SN1 mechanism, but this reaction is much less successful than with tertiary halogenoalkanes, because the secondary carbocation formed isn't as stable as a tertiary one.

Once the carbocation has been formed, it will react immediately with a cyanide ion. The lone pair on the cyanide ion is strongly attracted to the positive carbon, moves towards it, and forms a bond.

THE NUCLEOPHILIC SUBSTITUTION REACTIONS BETWEEN HALOGENOALKANES AND AMMONIA

This page gives you the facts and simple, uncluttered mechanisms for the nucleophilic substitution reactions between

halogenoalkanes and ammonia to produce primary amines. If you want the mechanisms explained to you in detail, there is a link at the bottom of the page. If you are interested in further substitution reactions, you will also find a link to a separate page dealing with these.

The reaction of primary halogenoalkanes with ammonia


Important! If you aren't sure about the difference between primary, secondary and tertiary halogenoalkanes, it is essential that you follow this link before you go on. Use the BACK button on your browser to return to this page.

The facts The halogenoalkane is heated with a concentrated solution of ammonia in ethanol. The reaction is carried out in a sealed tube. You couldn't heat this mixture under reflux, because the ammonia would simply escape up the condenser as a gas. We'll talk about the reaction using 1-bromoethane as a typical primary halogenoalkane. The reaction happens in two stages. In the first stage, a salt is formed - in this case, ethylammonium bromide. This is just like ammonium bromide, except that one of the hydrogens in the ammonium ion is replaced by an ethyl group.

There is then the possibility of a reversible reaction between this salt and excess ammonia in the mixture.

The ammonia removes a hydrogen ion from the ethylammonium ion to leave a primary amine - ethylamine. The more ammonia there is in the mixture, the more the forward reaction is favoured.

Note: You will find considerable disagreement in textbooks and other sources about the exact nature of the products in this reaction. Some of the information you'll come across is simply wrong! You can read the arguments about the products of this reaction by following this link.

The mechanism The mechanism involves two steps. The first is a simple nucleophilic substitution reaction:

Because the mechanism involves collision between two species in this slow step of the reaction, it is known as an SN2 reaction.
Note: Unless your syllabus specifically mentions SN2 by name, you can just call it nucleophilic substitution.

In the second step of the reaction an ammonia molecule may remove one of the hydrogens on the -NH3+. An ammonium ion is formed, together with a primary amine - in this case, ethylamine.

This reaction is, however, reversible. Your product will therefore contain a mixture of ethylammonium ions, ammonia, ethylamine and ammonium ions. Your major product will only be ethylamine if the ammonia is present in very large excess. Unfortunately the reaction doesn't stop here. Ethylamine is a good nucleophile, and goes on to attack unused bromoethane. This gets so complicated that it is dealt with on a separate page.

You will find a link at the bottom of this page.

The reaction of tertiary halogenoalkanes with ammonia


The facts The facts of the reactions are exactly the same as with primary halogenoalkanes. The halogenoalkane is heated in a sealed tube with a solution of ammonia in ethanol. For example:

Followed by:

The mechanism This mechanism involves an initial ionisation of the halogenoalkane:

followed by a very rapid attack by the ammonia on the carbocation (carbonium ion) formed:

This is again an example ofnucleophilic substitution. This time the slow step of the reaction only involves one species - the halogenoalkane. It is known as an SN1 reaction. There is a second stage exactly as with primary halogenoalkanes. An ammonia molecule removes a hydrogen

ion from the -NH3+ group in a reversible reaction. An ammonium ion is formed, together with an amine.

The reaction of secondary halogenoalkanes with ammonia


It is very unlikely that any of the current UK-based syllabuses for 16 - 18 year olds will ask you about this. In the extremely unlikely event that you will ever need it, secondary halogenoalkanes use both an SN2 mechanism and an SN1. Make sure you understand what happens with primary and tertiary halogenoalkanes, and then adapt it for secondary ones should ever need to.

EXPLAINING THE NUCLEOPHILIC SUBSTITUTION REACTIONS BETWEEN HALOGENOALKANES AND AMMONIA

This page guides you through the nucleophilic substitution mechanisms for the reactions between halogenoalkanes and ammonia to produce primary amines. It deals only with primary and tertiary halogenoalkanes. No current UK-based syllabus for 16 - 18 year olds is likely to ask about the reaction between secondary halogenoalkanes and ammonia. (It's not difficult - it's just not there!) If you are interested in further substitution reactions, at the bottom of this page you will find a link to a separate page dealing with these.
Important! It would help if you first read the page What is nucleophilic substitution? before you go on. You must also be clear about the differences between primary,

secondary and tertiary halogenoalkanes.

The reaction between primary halogenoalkanes and ammonia - the SN2 mechanism
Ammonia as a nucleophile A nucleophile is a species (an ion or a molecule) which is strongly attracted to a region of positive charge in something else. Nucleophiles are either fully negative ions, or else have a strongly - charge somewhere on a molecule. Ammonia obviously doesn't carry a negative charge. However, nitrogen is more electronegative than hydrogen, and so the nitrogen atom carries some degree of negative charge. It also has an active lone pair of electrons. The attack on the + carbon atom in the halogenoalkane is by this lone pair on the nitrogen atom.
Note: If you aren't sure about electronegativity follow this link before you read on. Use the BACK button on your browser to return to this page.

The nucleophilic substitution reaction - an SN2 reaction We'll talk this reaction through with bromoethane as a typical primary halogenoalkane. The bromoethane has a polar bond between the carbon and the bromine. The lone pair on the nitrogen will be strongly attracted to the + carbon, and will move towards it, beginning to form a bond with it. The approaching lone pair will repel the electrons in the carbon-bromine bond closer and closer to the bromine. The movement goes on until the ammonia is firmly attached to

the carbon, and the bromine has been expelled as a Br- ion.

Notice that the nitrogen in the product ion carries a positive charge (highlighted in red to draw attention to it). That charge has to be there for two reasons:

On the left hand side of the equation you start with two overall neutral molecules. Assuming you forgot about the positive charge, you would end up with a neutral species and a negative ion on the right. Charges must balance in equations, so something is wrong. The nitrogen looks wrong! The nitrogen atom is joined to 4 things rather than its usual 3. Nitrogen can only join to 4 things if it carries a positive charge.

Note: Nitrogen with a positive charge has the same arrangement of electrons as a carbon atom - which normally forms 4 bonds.

Technically, this is known as an SN2 reaction. S stands for substitution, N for nucleophilic, and the 2 is because the initial stage of the reaction involves two species - the bromoethane and the ammonia. If your syllabus doesn't refer to SN2 reactions by name, you can just call it nucleophilic substitution. The product of this reaction is a salt called ethylammonium bromide. Ethylammonium bromide is a salt of a primary amine and the acid, HBr. A primary amine has the formula R-NH2. It is primary in the sense that there is only one alkyl group attached to the nitrogen atom. Primary amines are weak bases very similar to ammonia and so form salts with acids. For example, ammonia reacts with HBr to give ammonium bromide, NH4+ Br-. The salt in the equation above is the one formed from ethylamine, CH3CH2NH2, and HBr.

At this point you need to know whether your examiners are happy for you to stop there, or want you to go a stage further to form the free primary amine.
Note: You need to check the mark schemes for recent exam papers, or any support material published by your examiners. You won't be able to tell this by looking at your syllabus, or at the exam papers themselves. You can find out how to get hold of this material by visiting your examiners' web site. If you are working to a UK-based syllabus for 16 18 year olds, you can find a link to this on the syllabuses page.

If they want the second step, an ammonia molecule now removes a hydrogen ion from the nitrogen.

The organic product is ethylamine, CH3CH2NH2. Notice that this change is reversible. The curly arrows for the reverse change haven't been shown to avoid confusion, but you can easily picture the lone pair on the nitrogen reclaiming its hydrogen from the NH4+ ion. What you will end up with is a mixture of all four of the species in this last equation - together, of course, with the bromide ions formed in the first stage. The higher the proportion of ammonia in the original reaction mixture, the greater the chance of the free ethylamine being formed.

The reaction between tertiary halogenoalkanes and ammonia - the SN1 mechanism
We'll talk this mechanism through with a simple tertiary halogenoalkane like the one on the right (2-bromo-2-methylpropane).

Why do tertiary halogenoalkanes need a different mechanism?

When a nucleophile attacks a primary halogenoalkane, it approaches the + carbon atom from the side away from the halogen atom. Any other approach is prevented by the halogen atom, which is both bulky and slightly negatively charged. The charge repels the incoming nucleophile. With a tertiary halogenoalkane, this approach from the back is impossible. The back of the molecule is completely cluttered with CH3 groups. The SN1 mechanism In the first stage, a small proportion of the halogenoalkane ionises to give a carbocation (carbonium ion) and a bromide ion.

This reaction is possible because tertiary carbocations are relatively stable compared with secondary or primary ones. Even so, the reaction is slow.
Note: Not sure about the stability of carbocations (carbonium ions)? Follow this link if you need to find out.

Once the carbocation is formed, however, it would react immediately it came into contact with an ammonia molecule. The lone pair on the nitrogen is strongly attracted towards the positive carbon, and moves towards it to create a new bond.

How fast the reaction happens overall is going to be governed by how fast the halogenoalkane ionises - because that's a slow process. Because this initial slow step only involves one species, the mechanism is described as SN1 - substitution, nucleophilic, one species taking part in the initial slow step. A salt is formed, and again you need to know whether your

examiners want you to go on beyond this to the formation of the free amine. If they do, one of the hydrogens attached to the nitrogen is removed by another ammonia molecule.

MULTIPLE NUCLEOPHILIC SUBSTITUTION IN THE REACTION BETWEEN HALOGENOALKANES AND AMMONIA

This page looks at further substitution in the nucleophilic substitution reaction between halogenoalkanes and ammonia following the formation of the primary amine. You could also think of this as a description of how primary, secondary and tertiary amines act as nucleophiles in a sequence of reactions with a halogenoalkane. Only primary halogenoalkanes are considered on this page. Only one current A' level syllabus (AQA) is likely to ask about these reactions, and that only asks about primary halogenoalkanes in this context.
Warning! Don't even think about reading this page unless you are confident about the nucleophilic substitution reaction involving a primary halogenoalkane and ammonia giving a primary amine. You need only worry about the SN2 reaction. The reactions on the current page follow on from those described on that page. They are no harder, but they do look more difficult. They are certainly much more tedious! It is important that you understand what is explained on this page. Life is too valuable to waste it learning each of the reactions described here. If you understand what's happening, you can work out the equations if you need to.

Why do you get multiple substitution?


The initial substitution - a reminder The lone pair on the nitrogen atom in an ammonia molecule is attracted towards the + carbon in the halogenoalkane - in this

example, bromoethane. It forms a bond with it - in the process expelling the bromine as a bromide ion.

A salt is formed - ethylammonium bromide. There is then the possibility of another ammonia molecule removing a hydrogen from the positive ion to give a primary amine - ethylamine. A primary amine has the general formula RNH2. This reaction is reversible, and you will only get significant amounts of the free amine if you use a large excess of ammonia.

The product as a nucleophile At the end of the initial substitution there will be a certain amount of free primary amine formed - the CH3CH2NH2 in the example above. There may not be very much of it, but there will be some. If that small amount reacts with something else, the reverse reaction can't happen any longer. The free amine will continue to be produced, and what was originally a reversible reaction becomes a one-way reaction - provided the free amine is removed by some other reaction as soon as it is formed. Now compare the structures of ammonia and ethylamine:

You can think of the primary amine as a slightly modified ammonia molecule. It has a lone pair on the nitrogen atom and an even bigger - charge than in ammonia.

That means that the primary amine is going to be a better nucleophile than ammonia is. You can therefore get a reaction between it and a molecule of the halogenoalkane.
Note: If you are interested in why the nitrogen in ethylamine has a greater degree of negative charge than the one in ammonia, you could follow this link and read about organic bases.

Making a secondary amine from a primary amine


What is a secondary amine? A secondary amine has the general formula R2NH. It is like an ammonia molecule (NH3) in which two of the hydrogens have been replaced by alkyl groups. The mechanism

The lone pair on the nitrogen in the primary amine attacks the + carbon exactly the same as the ammonia did. Bromine is lost as a bromide ion, and the immediate product is a salt called diethylammonium bromide - (CH3CH2)2NH2+ Br-. This is essentially ammonium bromide in which two of the hydrogens attached to the nitrogen have been replaced by ethyl groups. This then reacts with ammonia in a reversible reaction, exactly as we've seen before:

The organic product is a secondary amine - diethylamine. It is

secondary because there are two alkyl groups attached to the nitrogen atom.

Making a tertiary amine from the secondary amine


What is a tertiary amine? A tertiary amine has the general formula R3N. It is like an ammonia molecule in which all three of the hydrogens have been replaced by alkyl groups. The mechanism The secondary amine still has an active lone pair of electrons on the nitrogen atom. That, in turn, can attack bromoethane if it happens to collide with it. The two step sequence is exactly as before:

In this first step, the bromine is again displaced as a bromide ion and you get a salt formed called triethylammonium bromide.
Note: It has been necessary to rearrange the formula of the secondary amine in order to keep the appearance of the mechanism the same as the previous ones. Look carefully at it to be sure there isn't anything nasty going on!

An ammonia molecule can then remove the hydrogen from the nitrogen in the reversible reaction:

The organic product of this reaction is the tertiary amine, triethylamine. It is tertiary because of the three alkyl groups attached to the nitrogen.

The final stage - making a quaternary ammonium salt


What is a quaternary ammonium salt? A quaternary ammonium salt is an ammonium salt (for example, NH4+ Br-) in which all the hydrogens have been replaced by an alkyl group - for example, (CH3CH2)4N+ Br-. The mechanism The tertiary amine still has an active lone pair on the nitrogen and, once again, that can attack the + carbon in the bromoethane.

But this time there is nowhere else for the reaction to go. There is no longer a hydrogen atom on the nitrogen that an ammonia molecule could remove, and so the reaction finally comes to an end. The product is a salt called tetraethylammonium bromide, (CH3CH2)4N+ Br-.

If you react bromoethane and ammonia, what do you actually get?


You will always get a mixture of all the above products unless you use a very large excess of bromoethane so that there is enough of it to carry the sequence all the way through to the quaternary ammonium salt. You could favour formation of the primary amine (or its salt) by using a very large excess of ammonia. That way there is always a greater chance of a bromoethane molecule being hit by an

ammonia molecule than any other possible nucleophile. Even so, there will always, by chance, be some collisions leading to the follow-on products.

ELIMINATION MECHANISMS MENU


Elimination reactions involving halogenoalkanes
Elimination from 2-bromopropane . . . The formation of an alkene (propene) from 2-bromopropane. Elimination from unsymmetrical halogenoalkanes . . . Explains how to cope with cases where more than one elimination product can be formed from a single halogenoalkane. Elimination v. substitution . . . Discusses how the reaction between a halogenoalkane and hydroxide ions can lead to either an elimination reaction or nucleophilic substitution.

The dehydration of alcohols


The dehydration of ethanol . . . Facts and mechanism for the dehydration of ethanol using an acid catalyst to give ethene. The dehydration of more complicated alcohols . . . Explains how to cope with cases where more than dehydration product is possible.

THE ELIMINATION REACTIONS PRODUCING ALKENES FROM SIMPLE HALOGENOALKANES


This page gives you the facts and a simple, uncluttered mechanism for the elimination reaction between a simple halogenoalkane like 2-bromopropane and hydroxide ions (from, for example, sodium hydroxide) to give an alkene like propene. If you want the mechanism explained to you in detail, there is a link at the bottom of the page. You will also find a link to a page on elimination from more complicated halogenoalkanes where more than one product may be formed. Exam questions on this topic frequently ask about another possibility in the reactions between halogenoalkanes and hydroxide ions - nucleophilic substitution. There is also a link to a page discussing this.
Note: The competition between substitution and elimination (including the conditions needed and the mechanisms for both) is a rich source of exam questions if your syllabus includes it. You will probably find that the questions centre around secondary halogenoalkanes like 2-bromopropane, because these can easily be persuaded to do either reaction. That's why this page deals exclusively with 2-bromopropane.

The elimination reaction involving 2-bromopropane and hydroxide ions


The facts 2-bromopropane is heated under reflux with a concentrated solution of sodium or potassium hydroxide in ethanol. Heating under reflux involves heating with a condenser placed vertically in the flask to avoid loss of volatile liquids. Propene is formed

and, because this is a gas, it passes through the condenser and can be collected.

Everything else present (including anything formed in the alternative substitution reaction) will be trapped in the flask. The mechanism In elimination reactions, the hydroxide ion acts as a base removing a hydrogen as a hydrogen ion from the carbon atom next door to the one holding the bromine. The resulting re-arrangement of the electrons expels the bromine as a bromide ion and produces propene.

EXPLAINING THE ELIMINATION REACTIONS PRODUCING ALKENES FROM SIMPLE HALOGENOALKANES


This page guides you through the elimination mechanism for the reaction between simple halogenoalkanes like 2-bromopropane and hydroxide ions from, for example, sodium hydroxide. Elimination involving more complicated halogenoalkanes and the competition between elimination and substitution in these reactions are dealt with on separate pages. These are essential parts of this topic, and you should follow the links at the bottom of this page if you haven't already done so.

The elimination reaction involving 2-bromopropane and hydroxide ions

The role of the OH- ion in an elimination reaction Hydroxide ions have a very strong tendency to combine with hydrogen ions to make water - in other words, the OH- ion is a very strong base. In an elimination reaction, the hydroxide ion hits one of the hydrogen atoms in the CH3 group and pulls it off. This leads to a cascade of electron pair movements resulting in the formation of a carbon-carbon double bond, and the loss of the bromine as Br-.
Note: If you aren't happy about the use of curly arrows in mechanisms, follow this link before you go on. Use the BACK button on your browser to return to this page.

The complete elimination mechanism

The OH- ion takes one of the hydrogens from the CH3 group, but it only needs the hydrogen nucleus (a hydrogen ion). That means that the two electrons which originally joined the hydrogen to the carbon aren't being used any more. Those two electrons move to form a double bond between the two carbon atoms. The approach of those electrons repels the electrons in the carbon-bromine bond right out onto the bromine, throwing the bromine off as a negative ion.

The attack could equally well have been on any of the other hydrogens on the left-hand carbon, or on any on the right-hand one - it simply depends on what the OH- ion hit.
Taking chemistry further: The mechanism being described here is known as an E2

mechanism. The E stands for elimination, and the 2 is because the initial slow part of the reaction involves 2 species. This mechanism is used by primary and secondary halogenoalkanes. There is also an E1 mechanism which would apply to tertiary halogenoalkanes. (Secondary ones will also do it to some extent.) No current UK A level syllabus (or its equivalent) is likely to ask you about

ELIMINATION FROM UNSYMMETRIC HALOGENOALKANES


This page looks at elimination from unsymmetric halogenoalkanes such as 2bromobutane.

2-bromobutane is an unsymmetric halogenoalkane in the sense that it has a CH3 group one side of the C-Br bond and a CH2CH3 group the other. You have to be careful with compounds like this because of the possibility of more than one elimination product depending on where the hydrogen is removed from. The basic facts and mechanisms for these reactions are exactly the same as with simple halogenoalkanes like 2-bromopropane. This page only deals with the extra problems created by the possibility of more than one elimination product.
Important! What follow assumes that you are familiar with the mechanism for elimination from 2bromopropane. If you aren't, it is essential that you follow this link before you go on.

Background to the mechanism You will remember that elimination happens when a hydroxide ion (from, for example, sodium hydroxide) acts as a base and

removes a hydrogen as a hydrogen ion from the halogenoalkane. For example, in the simple case of elimination from 2bromopropane:

The hydroxide ion removes a hydrogen from one of the carbon atoms next door to the carbon-bromine bond, and the various electron shifts then lead to the formation of the alkene - in this case, propene. With an unsymmetric halogenoalkane like 2-bromobutane, there are several hydrogens which might possibly get removed. You need to think about each of these possibilities. Where does the hydrogen get removed from? The hydrogen has to be removed from a carbon atom adjacent to the carbon-bromine bond. If an OH- ion hit one of the hydrogens on the right-hand CH3 group in the 2-bromobutane (as we've drawn it), there's nowhere for the reaction to go.

To make room for the electron pair to form a double bond between the carbons, you would have to expel a hydrogen from the CH2 group as a hydride ion, H-. That is energetically much too difficult, and so this reaction doesn't happen. That still leaves the possibility of removing a hydrogen either

from the left-hand CH3 or from the CH2 group. If it was removed from the CH3 group:

The product is but-1-ene, CH2=CHCH2CH3. If it was removed from the CH2 group:

This time the product is but-2-ene, CH3CH=CHCH3. In fact the situation is even more complicated than it looks, because but-2-ene exhibits geometric isomerism. You get a mixture of two isomers formed - cis-but-2-ene and trans-but-2ene.

Cis-but-2-ene is also known as (Z)-but-2-ene; trans-but-2-ene is also known as (E)-but-2-ene. For an explanation of the two ways of naming these two compounds, follow the link in the box below. Which isomer gets formed is just a matter of chance.
Geometric isomerism: Isomerism is where you can draw more than one arrangement of the atoms

for a given molecular formula. Geometric isomerism is a special case of this involving molecules which have restricted rotation around one of the bonds - in this case, a carbon-carbon double bond. The C=C bond could only rotate if enough energy is put in to break the pi bond. Effectively, except at high temperatures, the C=C bond is "locked". In the case of but-2-ene, the two CH3 groups will either both be locked on one side of the C=C (to give the cis or (Z) isomer), or on opposite sides (to give the trans or (E) one). For a full discussion of geometric isomerism follow this link. Use the BACK button on your browser to return to this page. Beware! It is easy to miss geometric isomers in an exam. Always draw alkenes with the correct 120 bond angles around the C=C bond as shown in the diagrams for the cis and trans isomers above. If you take a short cut and write but-2-ene as CH3CH=CHCH3, you will almost certainly miss the fact that cis and trans forms are possible. This is a rich source of questions in an exam. You could easily throw away marks if you miss these possibilities.

The overall result Elimination from 2-bromobutane leads to a mixture containing:


but-1-ene cis-but-2-ene (also known as (Z)-but-2-ene) trans-but-2-ene (also known as (E)-but-2-ene)

ELIMINATION VERSUS SUBSTITUTION IN HALOGENOALKANES


This page discusses the factors that decide whether halogenoalkanes undergo elimination reactions or nucleophilic substitution when they react with hydroxide ions from, say,

sodium hydroxide or potassium hydroxide. Details for each of these types of reaction are given elsewhere, and you will find links to them from this page.

The reactions
Both reactions involve heating the halogenoalkane under reflux with sodium or potassium hydroxide solution. Nucleophilic substitution The hydroxide ions present are good nucleophiles, and one possibility is a replacement of the halogen atom by an -OH group to give an alcohol via a nucleophilic substitution reaction.

In the example, 2-bromopropane is converted into propan-2-ol.


Note: If you want to read about nucleophilic substitution in this reaction in detail, follow this link.

Elimination Halogenoalkanes also undergo elimination reactions in the presence of sodium or potassium hydroxide.

The 2-bromopropane has reacted to give an alkene - propene. Notice that a hydrogen atom has been removed from one of the end carbon atoms together with the bromine from the centre one. In all simple elimination reactions the things being removed are on adjacent carbon atoms, and a double bond is set up between those carbons.
Note: If you want to read about elimination in this

reaction in detail, follow this link.

What decides whether you get substitution or elimination?


The reagents you are using are the same for both substitution or elimination - the halogenoalkane and either sodium or potassium hydroxide solution. In all cases, you will get a mixture of both reactions happening - some substitution and some elimination. What you get most of depends on a number of factors. The type of halogenoalkane This is the most important factor. type of halogenoalkane substitution or elimination? primary secondary tertiary mainly substitution both substitution and elimination mainly elimination
Important! If you aren't clear about the various types of halogenoalkanes, it is essential that you follow this link before you read on. Use the BACK button on your browser to return to this page.

For example, whatever you do with tertiary halogenoalkanes, you will tend to get mainly the elimination reaction, whereas with primary ones you will tend to get mainly substitution. However, you can influence things to some extent by changing the conditions. The solvent

The proportion of water to ethanol in the solvent matters.


Water encourages substitution. Ethanol encourages elimination.

The temperature Higher temperatures encourage elimination. Concentration of the sodium or potassium hydroxide solution Higher concentrations favour elimination. In summary For a given halogenoalkane, to favour elimination rather than substitution, use:

heat a concentrated solution of sodium or potassium hydroxide pure ethanol as the solvent

Note: The explanations for these effects are well beyond the demands of UK A level syllabuses. Some things you just have to know!

The role of the hydroxide ions


The role of the hydroxide ion in a substitution reaction In the substitution reaction between a halogenoalkane and OH- ions, the hydroxide ions are acting as nucleophiles. For example, one of the lone pairs on the oxygen can attack the slightly positive carbon. This leads on to the loss of the bromine as a bromide ion, and the -OH group becoming attached in its place.

The role of the hydroxide ion in an elimination reaction Hydroxide ions have a very strong tendency to combine with hydrogen ions to make water - in other words, the OH- ion is a very strong base. In an elimination reaction, the hydroxide ion hits one of the hydrogen atoms in the CH3 group and pulls it off. This leads to a cascade of electron pair movements resulting in the formation of a carboncarbon double bond, and the loss of the bromine as Br-.
Note: The competition between substitution and elimination (including the conditions needed and the mechanisms for both) is a rich source of exam questions if your syllabus includes it. You will probably find that the questions centre around secondary halogenoalkanes like 2-bromopropane, because these can easily be persuaded to do either reaction. Important: If the elimination mechanism is on your syllabus, you are quite likely to be asked questions in which the substitution reaction crops up as well. If you need to revise this in detail, now is a good time to do it.

THE DEHYDRATION OF ALCOHOLS


This page gives you the facts and a simple, uncluttered mechanism for the acid catalysed dehydration of a simple alcohol like ethanol to give an alkene like ethene. If you want the mechanism explained to you in detail, there is a link at the bottom of the page. You will also find a link to a page on the dehydration of more complicated alcohols where more than one product may be formed.

The dehydration of ethanol


The facts

Ethanol can be dehydrated to give ethene by heating it with an excess of concentrated sulphuric acid at about 170C. Concentrated phosphoric(V) acid, H3PO4, can be used instead.

The acids aren't written into the equation because they serve as catalysts. If you like, you could write, for example, "conc H2SO4" over the top of the arrow.
Note: There are many side reactions which go on at the same time. These aren't required by any current A' level syllabus.

The mechanism - the full version We are going to discuss the mechanism using sulphuric acid. Afterwards, we'll describe how you can use a simplified version which will work for any acid, including phosphoric(V) acid. In the first stage, one of the lone pairs of electrons on the oxygen picks up a hydrogen ion from the sulphuric acid. The alcohol is said to be protonated.

The protonated ethanol loses a water molecule to give a carbocation (a carbonium ion).

Finally, a hydrogensulphate ion (from the sulphuric acid) pulls off a hydrogen ion from the carbocation.

The mechanism - a simplified version People normally quote a simplified version of this mechanism. The only Exam Board to want the mechanism for the dehydration of alcohols (AQA) is happy to accept this version. Instead of showing the full structure of the sulphuric acid, you write it as if it were simply a hydrogen ion, H+. That leaves the full mechanism:

An advantage of this (apart from the fact that it doesn't require you to draw the structure of sulphuric acid) is that it can be used for any acid catalyst without changing it at all. For example, if you use this version, you wouldn't need to worry about the structure of phosphoric(V) acid.
Note: Although most people probably write the mechanism in this form, it is actually quite misleading because it suggests the possibility of a free hydrogen ion in a chemical system. A free hydrogen ion is a raw proton and this is always attached to something else during a chemical reaction.

Personally, I find this simplification sloppy - but if your examiners are happy to accept it, who am I to argue! Go for the simple life!

EXPLAINING THE DEHYDRATION OF ALCOHOLS


This page guides you through the mechanism for the acid catalysed dehydration of a simple alcohol like ethanol to give an alkene like ethene. Dehydration of more complicated alcohols is dealt with on a separate page. This is an essential part of this topic, and you should follow the link at the bottom of this page if you haven't already done so.

The dehydration of ethanol


The full version of the mechanism This full version covers the mechanism using sulphuric acid. Afterwards, we'll look at the simplified version which will work for any acid, including phosphoric(V) acid. The oxygen atom in the ethanol has two active lone pairs of electrons, and one of these picks up a hydrogen ion from the sulphuric acid. The alcohol is said to be protonated.

Note: Only one of the lone pairs on the oxygen is being shown because only one is used in the reaction. Similarly, all the multitude of lone pairs in the sulphuric acid are omitted because they aren't relevant to the mechanism.

If you aren't happy about the use of curly arrows in mechanisms, follow this link before you go on. Use the BACK button on your browser to return to this page.

The negative ion produced is the hydrogensulphate ion, HSO4-. Notice that the oxygen atom in the alcohol has gained a positive charge. That charge has to be there for two reasons:

On the left hand side of the equation you start with two overall neutral molecules. Assuming you forgot about the positive charge, you would end up with a neutral species and a negative ion on the right. Charges must balance in equations, so something is wrong. The oxygen looks wrong! The oxygen atom is joined to 3 things rather than its usual 2. Oxygen can only join to 3 things if it carries a positive charge.

Note: Oxygen with a positive charge has the same arrangement of electrons as a nitrogen atom - which normally forms 3 bonds.

In the second stage of the reaction the protonated ethanol loses a water molecule to leave a carbocation (previously known as a carbonium ion) - an ion with a positive charge on a carbon atom. The carbon atom is positive because it has lost the electron that it originally contributed to the carbon-oxygen bond. Both of the electrons in that bond have moved onto the oxygen atom, neutralising the oxygen's charge.

One of the several things that can now happen to this carbocation is for it to lose a hydrogen ion from the CH3 group. This hydrogen ion is pulled off by a hydrogensulphate ion to regenerate the sulphuric acid catalyst.

Note: The other things that might happen to the carbocation lead to products like ethoxyethane and ethyl hydrogensulphate. The mechanisms for these aren't required by any current A' level syllabus.

The simplified version of the mechanism Instead of showing the full structure of the sulphuric acid, it is commonly written as if it were simply a hydrogen ion, H+. An advantage of this (apart from the fact that it doesn't require you to draw the structure of sulphuric acid) is that it can be used for any acid catalyst without changing it at all. For example, if you use this version, you wouldn't need to worry about the structure of phosphoric(V) acid. AQA (the only Exam Board to demand the dehydration mechanism at the moment) is happy to accept this version.

In the first stage, the ethanol gets protonated exactly as before the only difference is that you are writing H+ instead of the full structure of the sulphuric acid. The second stage is identical to the one in the full version of the mechanism. The final stage shows a hydrogen ion "falling off" the carbocation - rather than being pulled off. This is seriously misleading, but it's what the examiners want!

DEHYDRATION OF MORE COMPLICATED ALCOHOLS

This page builds on your understanding of the acid catalysed dehydration of alcohols. You have to be wary with more complicated alcohols in case there is the possibility of more than one alkene being formed. Butan-2-ol is a good example of this, with no less than three different alkenes being formed when it is dehydrated. Butan-2-ol is just an example to illustrate the problems. It is important that you understand it so that you can work out what will happen in similar cases. It would be quite impossible for you to learn what happens with every single alcohol you might be presented with. The basic facts and mechanisms for these reactions are exactly the same as with a simple alcohol like ethanol. This page only deals with the extra problems created by the possibility of more than one dehydration product.
Important! What follow assumes that you are familiar with the mechanism for the dehydration of ethanol. If you aren't, it is essential that you follow this link before you go on.

Background
To make the diagrams less cluttered, we'll use the simplified

version of the mechanism showing gain and loss of H+. Remember that the mechanism takes place in three stages:

The alcohol is protonated by the acid catalyst. The protonated alcohol loses a water molecule to give a carbocation (carbonium ion). The carbocation formed loses a hydrogen ion and forms a double bond.

So, in the case of the dehydration of a simple alcohol like ethanol:

The dehydration of butan-2-ol


The first two stages There is nothing new at all in these stages. In the first stage, the alcohol is protonated by picking up a hydrogen ion from the sulphuric acid.

In the second stage, the positive ion then sheds a water

molecule and produces a carbocation.

The complication arises in the next step. When the carbocation loses a hydrogen ion, where is it going to come from? Where does the hydrogen get removed from? So that a double bond can form, it will have to come from one of the carbons next door to the one with the positive charge. If a hydrogen ion is lost from the CH3 group

But-1-ene is formed. If a hydrogen ion is lost from the CH2 group

This time the product is but-2-ene, CH3CH=CHCH3. In fact the situation is even more complicated than it looks, because but-2-ene exhibits geometric isomerism. You get a mixture of two isomers formed - cis-but-2-ene and trans-but-2ene.

Cis-but-2-ene is also known as (Z)-but-2-ene; trans-but-2-ene is also known as (E)-but-2-ene. For an explanation of the two ways of naming these two compounds, follow the link in the box

below. Which isomer gets formed is just a matter of chance.


Geometric isomerism: Isomerism is where you can draw more than one arrangement of the atoms for a given molecular formula. Geometric isomerism is a special case of this involving molecules which have restricted rotation around one of the bonds - in this case, a carbon-carbon double bond. The C=C bond could only rotate if enough energy is put in to break the pi bond. Effectively, except at high temperatures, the C=C bond is "locked". In the case of but-2-ene, the two CH3 groups will either both be locked on one side of the C=C (to give the cis or (Z) isomer), or on opposite sides (to give the trans or (E) one). For a full discussion of geometric isomerism follow this link. Use the BACK button on your browser to return to this page. Beware! It is easy to miss geometric isomers in an exam. Always draw alkenes with the correct 120 bond angles around the C=C bond as shown in the diagrams for the cis and trans isomers above. If you take a short cut and write but-2-ene as CH3CH=CHCH3, you will almost certainly miss the fact that cis and trans forms are possible. This is a rich source of questions in an exam. You could easily throw away marks if you miss these possibilities.

The overall result Dehydration of butan-2-ol leads to a mixture containing:


but-1-ene cis-but-2-ene (also known as (Z)-but-2-ene) trans-but-2-ene (also known as (E)-but-2-ene)

NUCLEOPHILIC ADDITION MECHANISMS MENU


The addition of hydrogen cyanide to aldehydes and ketones . . . Facts and mechanism for the nucleophilic addition of hydrogen cyanide, HCN, to aldehydes and ketones. The reduction of aldehydes and ketones . . . Facts and a simplified mechanism for the reduction of aldehydes and ketones using sodium tetrahydridoborate, NaBH4.

THE NUCLEOPHILIC ADDITION OF HYDROGEN CYANIDE TO ALDEHYDES AND KETONES


This page gives you the facts and simple, uncluttered mechanisms for the nucleophilic addition reactions between carbonyl compounds (specifically aldehydes and ketones) and hydrogen cyanide, HCN. If you want the mechanisms explained to you in detail, there is a link at the bottom of the page. Aldehydes and ketones behave identically in their reaction with hydrogen cyanide, and so will be considered together - although equations and mechanisms will be given for both types of compounds for the sake of completeness.

The reaction of aldehydes and ketones with hydrogen cyanide


The facts Hydrogen cyanide adds across the carbon-oxygen double bond in aldehydes and ketones to produce compounds known as hydroxynitriles. For example, with ethanal (an aldehyde) you get 2-

hydroxypropanenitrile:

With propanone (a ketone) you get 2-hydroxy-2methylpropanenitrile:

Note: When you are naming these compounds, don't forget that the longest carbon chain has to include the carbon in the CN group. In both of the above examples, the longest carbon chain is 3 carbons - hence the "prop" in both names. The carbon with the nitrogen attached is always counted as the number 1 carbon in the chain.

The reaction isn't normally done using hydrogen cyanide itself, because this is an extremely poisonous gas. Instead, the aldehyde or ketone is mixed with a solution of sodium or potassium cyanide in water to which a little sulphuric acid has been added. The pH of the solution is adjusted to about 4 - 5, because this gives the fastest reaction. The solution will contain hydrogen cyanide (from the reaction between the sodium or potassium cyanide and the sulphuric acid), but still contains some free cyanide ions. This is important for the mechanism.
Note: If the reaction is done using hydrogen cyanide itself, a little sodium hydroxide solution is added to produce some cyanide ions from the weakly acidic HCN. Again the pH of the solution is adjusted to around pH 5 - in other words, the sodium hydroxide is not added to excess. The rate of the reaction falls if the pH is any higher. Whichever set of reagents you use, the reaction contains the same mixture of hydrogen cyanide and cyanide ions.

The mechanisms These are examples of nucleophilic addition. The carbon-oxygen double bond is highly polar, and the slightly positive carbon atom is attacked by the cyanide ion acting as a nucleophile.
Nucleophile: A species (molecule or ion) which attacks a positive site in something else. Nucleophiles are either fully negative ions or contain a fairly negative region somewhere in a molecule. All nucleophiles have at least one active lone pair of electrons. When you write mechanisms for reactions involving nucleophiles, you must show that lone pair.

The mechanism for the addition of HCN to propanone In the first stage, there is a nucleophilic attack by the cyanide ion on the slightly positive carbon atom.

The negative ion formed then picks up a hydrogen ion from somewhere - for example, from a hydrogen cyanide molecule.

The hydrogen ion could also come from the water or the H3O+ ions present in the slightly acidic solution. You don't need to remember all of these. One equation is perfectly adequate.
Note: The product molecule here has been drawn differently from the one in the equation further up this page. It has been rotated through 90. There is no reason why you can't do that if it makes the appearance of the mechanism easier to follow.

The mechanism for the addition of HCN to ethanal As before, the reaction starts with a nucleophilic attack by the cyanide ion on the slightly positive carbon atom.

It is completed by the addition of a hydrogen ion from, for example, a hydrogen cyanide molecule.

Note: Again, the product molecule looks different from the one in the equation further up this page. The central carbon atom still has the same four groups attached, but to make the mechanism easier to follow, they are simply arranged differently. That's not a problem - we're still talking about the same substance.

Optical isomerism in 2-hydroxypropanenitrile When 2-hydroxypropanenitrile is made in this last mechanism, it occurs as a racemic mixture - a 50/50 mixture of two optical isomers. It is possible that you might be exected to explain how this arises.
Note: You almost certainly won't be able to tell whether or not you need to know this from the syllabus. You need to refer to recent exam papers and mark schemes. If you haven't already got these, you can obtain them from your Exam Board via links on the syllabuses page.

Optical isomerism occurs in compounds which have four different groups attached to a single carbon atom. In this case,

the product molecule contains a CH3, a CN, an H and an OH all attached to the central carbon atom. The reason for the formation of equal amounts of two isomers lies in the way the ethanal gets attacked. Ethanal is a planar molecule, and attack by a cyanide ion will either be from above the plane of the molecule, or from below. There is an equal chance of either happening.

Attack from one side will lead to one of the two isomers, and attack from the other side will lead to the other.
Note: This is probably as much as you need to know for exam purposes, but a full explanation of this is given on the "talk through" page. Follow the link below.

All aldehydes will form a racemic mixture in this way. Unsymmetrical ketones will as well. (A ketone can be unsymmetrical in the sense that there is a different alkyl group either side of the carbonyl group.) What matters is that the product molecule must have four different groups attached to the carbon which was originally part of the carbon-oxygen double bond. EXPLAINING THE NUCLEOPHILIC ADDITION OF HYDROGEN CYANIDE TO ALDEHYDES AND KETONES This page explains the mechanism for the nucleophilic addition reaction between carbonyl compounds (specifically aldehydes and ketones) and hydrogen cyanide. It also looks in some detail at why a racemic mixture is formed when hydrogen cyanide reacts with an aldehyde like ethanal.

Background Bonding in the carbonyl group: the carbon-oxygen double bond Oxygen is far more electronegative than carbon and so has a strong tendency to pull electrons in a carbon-oxygen bond towards itself. One of the two pairs of electrons that make up a carbon-oxygen double bond is even more easily pulled towards the oxygen. That makes the carbon-oxygen double bond very highly polar.
Note: If you aren't sure about electronegativity and bond polarity follow this link before you go on. If you are interested in really understanding the bonding in the carbon-oxygen double bond you could explore it in detail by following this link. You don't need to know about this to understand the rest of this page. You might find that following this link will take you some time, because you will probably have to explore several pages of background material as well.

The cyanide ion as a nucleophile A nucleophile is a species (either a negatively charged ion or a negative region in a polar molecule) which is attracted to a positive site in another substance. All nucleophiles contain an active lone pair of electrons. The cyanide ion comes from hydrogen cyanide, which is a covalent molecule. Hydrogen cyanide is very weakly acidic, which means that it can lose a hydrogen ion - although not very easily.

Notice that when the hydrogen is lost, it leaves its electron behind on the carbon. That leaves a lone pair of electrons on that carbon, together with a negative charge. It's essential to realise that in the cyanide ion the active lone pair and the charge are on the carbon atom and not the nitrogen.
Note: There is also a lone pair on the nitrogen atom (not shown to avoid confusion), but that isn't important because

that's not where the negative charge is.

Explaining the conditions for the reaction Remember that the reaction is done by reacting the aldehyde or ketone with a solution of sodium or potassium cyanide to which enough dilute sulphuric acid is added to give a pH of around 4 5. Hydrogen cyanide as a very weak acid The initial attack on the carbonyl group is by a cyanide ion. Although the reaction overall adds hydrogen cyanide across the double bond, using hydrogen cyanide as the reagent isn't successful because hydrogen cyanide is such a weak acid. It produces very, very few hydrogen ions and cyanide ions in solution. The number of cyanide ions available to attack the slightly positive carbon is extremely small and so the reaction would be very slow. Why is the potassium cyanide acidified slightly? The presence of an acid in solution helps to strengthen the polarity of the carbon-oxygen double bond. The electrons in that bond are strongly attracted towards the hydrogen ions in the solution.

Why isn't a lot more acid added to give a really low pH? You might think that the more acid you added, the more you could increase the polarity of the carbon-oxygen double bond. Unfortunately, there is a competing effect. The more acid you add, the more the cyanide ions get converted into hydrogen cyanide. Since cyanide ions are what actually

attacks the slightly positive carbon, removing them isn't helpful! A pH of 4 - 5 is found experimentally to give you the best rate of reaction. It increases the polarity of the double bond by a useful amount, but without removing too many of the cyanide ions as HCN.

The mechanisms
Mechanisms for the addition of hydrogen cyanide to aldehydes and ketones are given separately for completeness, but there is absolutely no difference between them. They are examples of nucleophilic addition. The mechanism for the addition of HCN to propanone As the cyanide ion approaches the slightly positive carbon atom, the lone pair of electrons is attracted towards the carbon and forms a bond with it. At the same time the two electrons in one of the bonds joining the carbon to the oxygen are repelled until they end up entirely on the oxygen - giving it a negative charge.

Note: If you aren't happy about the use of curly arrows in mechanisms, follow this link before you go on. Use the BACK button on your browser to return to this page.

The negative ion formed then picks up a hydrogen ion from somewhere - for example, from a hydrogen cyanide molecule.

The hydrogen ion could also come from the water or the H3O+ ions (often written as H+(aq)) present in the slightly acidic solution. You don't need to remember all of these. One equation is perfectly adequate. In fact, most sources show this final stage

as a reaction with just H+.

Note: Although most people (including examiners) probably take this short cut, it doesn't mean that it's right. It is sloppy because it suggests the possibility of free hydrogen ions in a chemical reaction - hydrogen ions are simply protons, and are always attached to something else. It also denies you the satisfaction of writing an equation which shows the production of a new cyanide ion which can go on to react with another molecule of the carbonyl compound. If you can't be bothered to do it properly, then do at least write + + the hydrogen ion as H (aq) - not just as H .

The mechanism for the addition of HCN to ethanal Exactly as before, as the cyanide ion approaches the slightly positive carbon atom, the lone pair of electrons is attracted towards the carbon and forms a bond with it. At the same time the two electrons in one of the bonds joining the carbon to the oxygen are repelled until they end up entirely on the oxygen giving it a negative charge.

The negative ion formed then picks up a hydrogen ion from somewhere - for example, from a hydrogen cyanide molecule.

Once again, most sources show this final stage as a reaction with just H+. If you must do it that way, then write the hydrogen ion as H+(aq) - not just as H+.

Optical isomerism in 2-hydroxypropanenitrile


2-hydroxypropanenitrile is the name of the product when ethanal reacts with hydrogen cyanide. It is formed in this reaction as an exactly equal mixture of two optical isomers, known as a racemic mixture.
Note: For a full discussion of optical isomerism follow this link. Use the BACK button on your browser to return to this page.

Optical isomerism occurs in compounds which have four different groups attached to a single carbon atom. In this case, the product molecule contains a CH3, a CN, an H and an OH all attached to the central carbon atom. The reason for the formation of equal amounts of two isomers lies in the way the ethanal gets attacked. Ethanal is a planar molecule, and attack by a cyanide ion will either be from above the plane of the molecule, or from below. There is an equal chance of either happening.

Attack from above will lead to one of the two isomers, and attack from below will lead to the other. The next diagram shows what happens if attack is from above the plane of the molecule. Notice that the existing groups get forced down away from the approaching cyanide ion. The diagram shows the final product after it has gained a hydrogen ion.

Attack from below forces the existing groups upwards.

Why are these two product molecules different from each other? Everything seems to be attached in the same way, but look what happens if you rotate the second molecule in space so that the cyanide group is at the top.

Now compare that with the molecule formed by attack from above.

They aren't the same! Although the CN and H are lined up the same way, the CH3 and OH are reversed. There is no way that you can rotate one molecule in space to make it look the same as the other one. These are therefore isomers. The relationship between them is that they are mirror images of

each other. If one isomer were to look in a mirror, what it would see would be the other one. Optical isomers are described as "non-superimposable mirror images". Because there is an equal chance of the attack coming from above or below the plane of the molecule, then you will get equal amounts of the two isomers formed - a racemic mixture. This argument applies to all aldehydes, and to ketones as long as they are unsymmetric - with a different alkyl group either side of the carbonyl group. A symmetric ketone like propanone, CH3COCH3, will only produce a single product - not a mixture of isomers. The product doesn't have four different groups around the central carbon atom, and so won't have optical isomers. If you followed the above argument through with propanone, you would find that you ended up with two molecules which could be rotated in space so that they were identical. Convince yourself by trying it! THE REDUCTION OF ALDEHYDES AND KETONES

This page gives you the facts and mechanisms for the reduction of carbonyl compounds (specifically aldehydes and ketones) using sodium tetrahydridoborate (sodium borohydride) as the reducing agent. Only one UK A level Exam Board (AQA) is likely to ask for these mechanisms, and they are happy with a simplified version of what is quite a complex mechanism. Because of that simplification, these reactions are dealt with entirely on this page - without the "talk through" page that you will find for other mechanisms on this site.

The reduction of aldehydes and ketones by sodium tetrahydridoborate The facts Sodium tetrahydridoborate (previously known as sodium borohydride) has the formula NaBH4, and contains the BH4-

ion. That ion acts as the reducing agent. There are several quite different ways of carrying out this reaction. Two possible variants (there are several others!) are:

The reaction is carried out in solution in water to which some sodium hydroxide has been added to make it alkaline. The reaction produces an intermediate which is converted into the final product by addition of a dilute acid like sulphuric acid. The reaction is carried out in solution in an alcohol like methanol, ethanol or propan-2-ol. This produces an intermediate which can be converted into the final product by boiling it with water.

In each case, reduction essentially involves the addition of a hydrogen atom to each end of the carbon-oxygen double bond to form an alcohol. Reduction of aldehydes and ketones lead to two different sorts of alcohol. The reduction of an aldehyde For example, with ethanal you get ethanol:

Notice that this is a simplified equation - perfectly acceptable to examiners. The H in square brackets means "hydrogen from a reducing agent". In general terms, reduction of an aldehyde leads to a primary alcohol. A primary alcohol is one which only has one alkyl group attached to the carbon with the -OH group on it. They all contain the grouping -CH2OH.
Note: There is one exception to this. Methanol CH3OH is also a primary alcohol. Think of this as H-CH2OH.

The reduction of a ketone

For example, with propanone you get propan-2-ol:

Reduction of a ketone leads to a secondary alcohol. A secondary alcohol is one which has two alkyl groups attached to the carbon with the -OH group on it. They all contain the grouping -CHOH.

Beware! The following mechanisms are simplified for UK A level purposes to the point that they are wrong! If you are working outside the UK A level system, please don't read any further!

The simplified mechanisms The BH4- ion is essentially a source of hydride ions, H-. The simplification used is to write H- instead of BH4-. Doing this not only makes the initial attack easier to write, but avoids you getting involved with some quite complicated boron compounds that are formed as intermediates. The reduction is an example of nucleophilic addition. The carbon-oxygen double bond is highly polar, and the slightly positive carbon atom is attacked by the hydride ion acting as a nucleophile. A hydride ion is a hydrogen atom with an extra electron - hence the lone pair.
Nucleophile: A species (molecule or ion) which attacks a positive site in something else. Nucleophiles are either fully negative ions or contain a fairly negative region somewhere in a molecule. All nucleophiles have at least one active lone pair of electrons. When you write mechanisms for reactions involving nucleophiles, you must show that lone pair.

The mechanism for the reduction of ethanal

In the first stage, there is a nucleophilic attack by the hydride ion on the slightly positive carbon atom. The lone pair of electrons on the hydride ion forms a bond with the carbon, and the electrons in one of the carbon-oxygen bonds are repelled entirely onto the oxygen, giving it a negative charge.

What happens now depends on whether you add an acid or water to complete the reaction. Adding an acid: When the acid is added, the negative ion formed picks up a hydrogen ion to give an alcohol.

Note: You may find that other sources write the hydrogen ion + simply as H . That's not good practice, because it suggests a free hydrogen ion. The hydrogen ion is actually attached to a water + molecule as H3O . Writing that makes the equation look more + complicated. H (aq) is a happy compromise.

Adding water: This time, the negative ion takes a hydrogen ion from a water molecule.

The mechanism for the reduction of propanone As before, the reaction starts with a nucleophilic attack by the hydride ion on the slightly positive carbon atom.

Again, what happens next depends on whether you add an acid or water to complete the reaction. Adding an acid: The negative ion reacts with a hydrogen ion from the acid added in the second stage of the reaction.

Adding water: This time, the negative ion takes a hydrogen ion from a water molecule.

Important! Remember that the equations and mechanisms given on this page are not the truth - they are merely simplifications to suit the demands of a particular A level syllabus.

NUCLEOPHILIC ADDITION / ELIMINATION MECHANISMS MENU

What is nucleophilic addition / elimination? . . . Includes an explanation of all the terms involved, together with a general mechanism for these reactions. It would be useful to read this before you look at specific examples. The reaction of acyl chlorides with water . . . Facts and mechanism for the formation of acids from acyl chlorides by reaction with water. The reaction of acyl chlorides with alcohols . . . Facts and mechanism for the formation of esters from acyl chlorides by reaction with alcohols. The reaction of acyl chlorides with ammonia . . . Facts and mechanism for the formation of amides from acyl chlorides by reaction with ammonia. The reaction of acyl chlorides with primary amines . . . Facts and mechanism for the formation of N-substituted amides from acyl chlorides by reaction with primary amines.

NUCLEOPHILIC ADDITION / ELIMINATION IN THE REACTIONS OF ACYL CHLORIDES


Background
What are acyl chlorides? Acyl chlorides (also known as acid chlorides) are one of a number of types of compounds known as "acid derivatives". This is ethanoic acid:

If you remove the -OH group and replace it by a -Cl, you have produced an acyl chloride.

This molecule is known as ethanoyl chloride and for the rest of this topic will be taken as typical of acyl chlorides in general. Acyl chlorides are extremely reactive. They are open to attack by nucleophiles - with the overall result being a replacement of the chlorine by something else. Nucleophiles A nucleophile is a species (an ion or a molecule) which is strongly attracted to a region of positive charge in something else. Nucleophiles are either fully negative ions, or else have a strongly - charge somewhere on a molecule. The nucleophiles that we shall be looking at all depend on lone pairs on either an oxygen atom or a nitrogen atom. Nucleophiles based on oxygen atoms We shall be talking about water and alcohols, taking ethanol as a typical alcohol.

Notice how similar these two molecules are around the oxygen atom. That's what turns out to be important. Nucleophiles based on nitrogen atoms We shall be considering ammonia and primary amines, taking

ethylamine as a typical primary amine. A primary amine contains the -NH2 group attached to either an alkyl group (as it is here) or a benzene ring. As far as these reactions are concerned, the nature of any hydrocarbon attached to the nitrogen makes no difference. The nitrogen atom is the important bit.

Again, notice how similar these two molecules are around the nitrogen atom - and also how similar they are to the previous ones containing oxygen. Both types of molecule have an active lone pair of electrons attached to one of the most electronegative elements. All of these molecules also have at least one hydrogen atom attached to the oxygen or nitrogen. Why are acyl chlorides attacked by nucleophiles? The carbon atom in the -COCl group has both an oxygen atom and a chlorine atom attached to it. Both of these are very electronegative. They both pull electrons towards themselves, leaving the carbon atom quite positively charged.

Note: If you aren't sure about electronegativity and bond polarity follow this link before you go on. Use the BACK button on your browser to return to this page.

The overall reaction We are going to generalise this for the moment by writing the reacting molecule as "Nu-H". Nu is the bit of the molecule which contains the nucleophilic oxygen or nitrogen atom. The attached hydrogen turns out to be essential to the reaction.

The general equation for the reaction is:

In each case, the net effect is that you replace the -Cl by -Nu, and hydrogen chloride is formed as well. Since the initial attack is by a nucleophile, and the overall result is substitution, it would seem reasonable to describe the reaction as nucleophilic substitution. However, the reaction happens in two distinct stages. The first involves an addition reaction, which is followed by an elimination reaction where HCl is produced. So the mechanism is also known as nucleophilic addition / elimination.
Note: You will find both terms in use - and to confuse the issue still further, these are also examples of condensation reactions. A condensation reaction is one in which two molecules join together to make a bigger one, and in the process shed a little molecule. The only Exam Board to require these mechanisms (AQA) calls them addition / elimination reactions.

The general mechanism


The addition stage of the mechanism As the lone pair on the nucleophile approaches the fairly positive carbon in the acyl chloride, it moves to form a bond with it. In the process, the two electrons in one of the carbon-oxygen bonds are repelled entirely onto the oxygen, leaving that oxygen negatively charged.

Note: If you aren't happy about the use of curly arrows in mechanisms, follow this link before you go on. Use the BACK button on your browser to return to this page.

Notice the positive charge that forms on the nucleophile. Just accept this for the moment. It's much easier to explain why that charge must be there if you have a real example in front of you. This is fully explained in the pages on the reactions involving water, ammonia and so on. The elimination stage of the mechanism This happens in two steps. In the first step, the carbon-oxygen double bond reforms. To make room for it, the electrons in the carbon-chlorine bond are repelled until they are entirely on the chlorine atom - forming a chloride ion.

Finally, the chloride ion plucks the hydrogen off the original nucleophile. It removes it as a hydrogen ion, leaving the pair of electrons behind on the oxygen or nitrogen atom in that nucleophile. That cancels the positive charge.

Important! Take some trouble to understand this general mechanism. When you read the pages on the specific reactions you should then be able to see all the similarities which exist - and realise that they are all essentially the same mechanism. If you try to learn them all as separate mechanisms, you are in for a lot of tedium!

NUCLEOPHILIC ADDITION / ELIMINATION IN THE REACTION BETWEEN ACYL CHLORIDES AND WATER
This page gives you the facts and a simple, uncluttered mechanism for the nucleophilic addition / elimination reaction between acyl chlorides (acid chlorides) and water. If you want the mechanism explained to you in detail, there is a link at the bottom of the page. Ethanoyl chloride is taken as a typical acyl chloride. Any other acyl chloride will behave in the same way. Simply replace the CH3 group in what follows by anything else you want.

The reaction between ethanoyl chloride and water


The facts Ethanoyl chloride reacts instantly with cold water. There is a very exothermic reaction in which a steamy acidic gas is given off (hydrogen chloride) and ethanoic acid is formed.

The mechanism The first stage (the addition stage of the reaction) involves a nucleophilic attack on the fairly positive carbon atom by one of the lone pairs on the oxygen of a water molecule.

Note: Only one of the two lone pairs on the oxygen in water is shown. This is to avoid cluttering an already complicated diagram with things that aren't relevant.

The second stage (the elimination stage) happens in two steps. In the first, the carbon-oxygen double bond reforms and a chloride ion is pushed off.

That is followed by removal of a hydrogen ion by the chloride ion to give ethanoic acid and hydrogen chloride.

EXPLAINING NUCLEOPHILIC ADDITION / ELIMINATION IN THE REACTION BETWEEN ACYL CHLORIDES AND WATER

This page guides you through the mechanism for the nucleophilic addition / elimination reaction between acyl chlorides (acid chlorides) and water. Ethanoyl chloride is taken as a typical acyl chloride. Any other

acyl chloride will behave in the same way. Simply replace the CH3 group in what follows by anything else you want.
Important! If you haven't already done so, it would help if you first read the page What is nucleophilic addition / elimination? before you go on.

The reaction between ethanoyl chloride and water


Water as a nucleophile A nucleophile is a species (an ion or a molecule) which is strongly attracted to a region of positive charge in something else. Nucleophiles are either fully negative ions, or else have a strongly - charge somewhere on a molecule. Oxygen is much more electronegative than hydrogen and so drags the electrons in the oxygen-hydrogen bonds towards itself. That produces a significant amount of negative charge on the oxygen atom. The oxygen also has two active lone pairs of electrons. One of these attacks the ethanoyl chloride.
Note: If you aren't sure about electronegativity and bond polarity follow this link before you go on. Use the BACK button on your browser to return to this page.

Why are acyl chlorides attacked by nucleophiles? The carbon atom in the -COCl group has both an oxygen atom and a chlorine atom attached to it. Both of these are very electronegative. They both pull electrons towards themselves, leaving the carbon atom quite positively charged. It is that carbon atom which is attacked by one of the lone pairs on the oxygen atom in a water molecule. The ethanoyl chloride molecule is also planar (flat) around that carbon atom, and that leaves plenty of room for a nucleophile to

attack either from above or below the plane of the molecule.

The mechanism The reaction happens in two main stages - an addition stage, followed by an elimination stage. In the addition stage, a water molecule becomes attached to the carbon in the ethanoyl chloride. As the lone pair on the water approaches the fairly positive carbon in the ethanoyl chloride, it moves to form a bond with it. In the process, the two electrons in one of the carbon-oxygen bonds are repelled entirely onto the oxygen, leaving that oxygen negatively charged.

Note: Only one of the two lone pairs on the oxygen in water is shown because only one of them is involved in the reaction. There's no reason why you couldn't write both in if you want to - it just clutters the diagram.

Notice that the oxygen atom in the water has gained a positive charge. The underlying reason for this is that when the lone pair forms a bond with the carbon, electrons are moving away from the oxygen. What matters, though, is that you remember to show the positive charge in an exam. Think of it like this: If you leave out the positive charge, two things are wrong with

the equation:

The charges don't balance. You start with two overall neutral molecules and, if you forgot the positive charge, you would end up with a negative ion. There has got to be a positive charge somewhere to balance the negative one. The oxygen looks wrong! Oxygen normally forms two bonds, but here it is forming three. Oxygen can only form three bonds if it carries a positive charge. (A positively charged oxygen atom has the same electronic structure as nitrogen - which normally forms three bonds.)

You can put both things right with a positive charge on the oxygen.

The elimination stage stage of the reaction happens in two steps. In the first, the carbon-oxygen double bond reforms. As the electron pair moves back to form a bond with the carbon, the pair of electrons in the carbon-chlorine bond are forced entirely onto the chlorine to give a chloride ion.

Finally, the chloride ion forms a bond with one of the hydrogens on the oxygen - removing it as a hydrogen ion. The electrons in the hydrogen-oxgyen bond move back onto the oxygen, cancelling the positive charge.

NUCLEOPHILIC ADDITION / ELIMINATION IN THE

REACTION BETWEEN ACYL CHLORIDES AND ALCOHOLS

This page gives you the facts and a simple, uncluttered mechanism for the nucleophilic addition / elimination reaction between acyl chlorides (acid chlorides) and alcohols. If you want the mechanism explained to you in detail, there is a link at the bottom of the page. Ethanoyl chloride is taken as a typical acyl chloride. Any other acyl chloride will behave in the same way. Simply replace the CH3 group in what follows by anything else you want. Similarly, ethanol is taken as a typical alcohol. If you are interested in another alcohol, you can replace the CH3CH2 group by any other alkyl group. The reaction between ethanoyl chloride and ethanol The facts Ethanoyl chloride reacts instantly with cold ethanol. There is a very exothermic reaction in which a steamy acidic gas is given off (hydrogen chloride). Ethyl ethanoate (an ester) is formed.

The mechanism The first stage (the addition stage of the reaction) involves a nucleophilic attack on the fairly positive carbon atom by one of the lone pairs on the oxygen of an ethanol molecule.

Note: Only one of the two lone pairs on the oxygen in the ethanol is shown. This is to avoid cluttering an already complicated diagram with things that aren't relevant.

The second stage (the elimination stage) happens in two steps. In the first, the carbon-oxygen double bond reforms and a chloride ion is pushed off.

That is followed by removal of a hydrogen ion by the chloride ion to give ethyl ethanoate and hydrogen chloride.

EXPLAINING NUCLEOPHILIC ADDITION / ELIMINATION IN THE REACTION BETWEEN ACYL CHLORIDES AND ALCOHOLS

This page guides you through the mechanism for the nucleophilic addition / elimination reaction between acyl chlorides (acid chlorides) and alcohols. Ethanoyl chloride is taken as a typical acyl chloride. Any other acyl chloride will behave in the same way. Ethanol is taken as a typical alcohol. Again, any other alcohol will behave in the same way.
Important! If you haven't already done so, it would help if

you first read the page What is nucleophilic addition / elimination? before you go on. The reaction between ethanoyl chloride and ethanol has a mechanism which is identical to that between ethanoyl chloride and water, but it looks more complicated. It would be a good idea to be sure about the water mechanism before you attack this one!

The reaction between ethanoyl chloride and ethanol


Ethanol as a nucleophile A nucleophile is a species (an ion or a molecule) which is strongly attracted to a region of positive charge in something else. Nucleophiles are either fully negative ions, or else have a strongly - charge somewhere on a molecule. Oxygen is much more electronegative than hydrogen and carbon and so drags bonding electrons towards itself. That produces a significant amount of negative charge on the oxygen atom. The oxygen also has two active lone pairs of electrons. One of these attacks the ethanoyl chloride. Why are acyl chlorides attacked by nucleophiles? The carbon atom in the -COCl group has both an oxygen atom and a chlorine atom attached to it. Both of these are very electronegative. They both pull electrons towards themselves, leaving the carbon atom quite positively charged. It is that carbon atom which is attacked by one of the lone pairs on the oxygen atom in an ethanol molecule. The ethanoyl chloride molecule is also planar (flat) around that carbon atom, and that leaves plenty of room for a nucleophile to attack either from above or below the plane of the molecule.

The mechanism The reaction happens in two main stages - an addition stage, followed by an elimination stage. In the addition stage, an ethanol molecule becomes attached to the carbon in the ethanoyl chloride. As the lone pair on the oxygen approaches the fairly positive carbon in the ethanoyl chloride, it moves to form a bond with it. In the process, the two electrons in one of the carbon-oxygen bonds are repelled entirely onto that oxygen, leaving it negatively charged.

Note: Only one of the two lone pairs on the oxygen in the ethanol is shown because only one of them is involved in the reaction. There's no reason why you couldn't write both in if you want to - it just clutters the diagram.

Notice that the oxygen atom in the ethanol has gained a positive charge. The underlying reason for this is that when the lone pair forms a bond with the carbon, electrons are moving away from the oxygen. What matters, though, is that you remember to show the positive charge in an exam. Think of it like this: If you leave out the positive charge, two things are wrong with

the equation:

The charges don't balance. You start with two overall neutral molecules and, if you forgot the positive charge, you would end up with a negative ion. There has got to be a positive charge somewhere to balance the negative one. The oxygen looks wrong! Oxygen normally forms two bonds, but here it is forming three. Oxygen can only form three bonds if it carries a positive charge. (A positively charged oxygen atom has the same electronic structure as nitrogen - which normally forms three bonds.)

You can put both things right with a positive charge on the oxygen.

The elimination stage stage of the reaction happens in two steps. In the first, the carbon-oxygen double bond reforms. As the electron pair moves back to form a bond with the carbon, the pair of electrons in the carbon-chlorine bond are forced entirely onto the chlorine to give a chloride ion.

Finally, the chloride ion forms a bond with the hydrogen on the oxygen - removing it as a hydrogen ion. The electrons in the hydrogen-oxgyen bond move back onto the oxygen, cancelling the positive charge.

NUCLEOPHILIC ADDITION / ELIMINATION IN THE REACTION BETWEEN ACYL CHLORIDES AND AMMONIA
This page gives you the facts and a simple, uncluttered mechanism for the nucleophilic addition / elimination reaction between acyl chlorides (acid chlorides) and ammonia. If you want the mechanism explained to you in detail, there is a link at the bottom of the page. Ethanoyl chloride is taken as a typical acyl chloride. Any other acyl chloride will behave in the same way. Simply replace the CH3 group in what follows by anything else you want.

The reaction between ethanoyl chloride and ammonia


The facts Ethanoyl chloride reacts violently with a cold concentrated solution of ammonia. A white solid product is formed which is a mixture of ethanamide (an amide) and ammonium chloride.

Notice that, unlike the reactions between ethanoyl chloride and water or ethanol, hydrogen chloride isn't produced - at least, not in any quantity. Any hydrogen chloride formed would immediately react with excess ammonia to give ammonium chloride.

The mechanism The first stage (the addition stage of the reaction) involves a nucleophilic attack on the fairly positive carbon atom by the lone pair on the nitrogen atom in the ammonia.

The second stage (the elimination stage) happens in two steps. In the first, the carbonoxygen double bond reforms and a chloride ion is pushed off.

That is followed by removal of a hydrogen ion from the nitrogen. This might happen in one of two ways: It might be removed by a chloride ion, producing HCl (which would immediately react with excess ammonia to give ammonium chloride as above) . . .

and

. . . or it might be removed directly by an ammonia molecule.

The ammonium ion, together with the chloride ion already there, makes up the ammonium chloride formed in the reaction.

EXPLAINING NUCLEOPHILIC ADDITION / ELIMINATION IN THE REACTION BETWEEN ACYL CHLORIDES AND AMMONIA
This page guides you through the mechanism for the nucleophilic addition / elimination reaction between acyl chlorides (acid chlorides) and ammonia. Ethanoyl chloride is taken as a typical acyl chloride. Any other acyl chloride will behave in the same way.
Important! If you haven't already done so, it would help if you first read the page What is nucleophilic addition / elimination? before you go on.

The reaction between ethanoyl chloride and ammonia


Ammonia as a nucleophile A nucleophile is a species (an ion or a molecule) which is strongly attracted to a region of positive charge in something else. Nucleophiles are either fully negative ions, or else have a strongly - charge somewhere on a molecule. Nitrogen is more electronegative than hydrogen and so drags the bonding electrons towards itself. That produces a significant amount of negative charge on the nitrogen atom. The nitrogen also has an active lone pair of electrons. It is this which attacks the ethanoyl chloride. Why are acyl chlorides attacked by nucleophiles? The carbon atom in the -COCl group has both an oxygen atom and a chlorine atom attached to it. Both of these are very electronegative. They both pull electrons towards themselves, leaving the carbon atom quite positively charged. It is that carbon atom which is attacked by the lone pair on the nitrogen atom in an ammonia molecule. The ethanoyl chloride molecule is also planar (flat) around that carbon atom, and that leaves plenty of room for a nucleophile to attack either from above or below the plane of the molecule.

The mechanism The reaction happens in two main stages - an addition stage, followed by an elimination stage. In the addition stage, an ammonia molecule becomes attached

to the carbon in the ethanoyl chloride. As the lone pair on the nitrogen approaches the fairly positive carbon in the ethanoyl chloride, it moves to form a bond with it. In the process, the two electrons in one of the carbon-oxygen bonds are repelled entirely onto the oxygen, leaving it negatively charged.

Notice that the nitrogen atom has gained a positive charge. The underlying reason for this is that when the lone pair forms a bond with the carbon, electrons are moving away from the nitrogen. What matters, though, is that you remember to show the positive charge in an exam. Think of it like this: If you leave out the positive charge, two things are wrong with the equation:

The charges don't balance. You start with two overall neutral molecules and, if you forgot the positive charge, you would end up with a negative ion. There has got to be a positive charge somewhere to balance the negative one. The nitrogen looks wrong! Nitrogen normally forms three bonds, but here it is forming four. Nitrogen can only form four bonds if it carries a positive charge. (A positively charged nitrogen atom has the same electronic structure as carbon - which normally forms four bonds.)

You can put both things right with a positive charge on the nitrogen.

The elimination stage stage of the reaction happens in two steps. In the first, the carbon-oxygen double bond reforms. As the electron pair moves back to form a bond with the carbon, the pair of electrons in the carbon-chlorine bond are forced entirely

onto the chlorine to give a chloride ion.

Finally, one of the hydrogens attached to the nitrogen is removed as a hydrogen ion. The electrons in the hydrogennitrogen bond move back onto the nitrogen, cancelling the positive charge. The hydrogen ion might be removed in one of two ways. The first way is entirely consistent with what happens in the reactions between water or ethanol and acyl chlorides. The hydrogen is removed by the chloride ion.

The hydrogen chloride produced would at once react with any excess ammonia present to form ammonium chloride.

The other (more likely) possibility is that the hydrogen ion gets removed directly by an ammonia molecule. This forms an ammonium ion. The ammonium ion together with the chloride ion formed in the previous stage makes up the ammonium chloride produced in the reaction.

Note: You probably don't really need to know both of these routes for the removal of the hydrogen ion. Choose for yourself. It doesn't matter very much, because both will happen. The route involving the formation of hydrogen chloride is exactly in line with the water or alcohol mechanisms, but you do need to remember to react the HCl with excess ammonia. The removal by another ammonia molecule is probably the major route as long as the ammonia is in excess which it almost certainly will be. The reaction is normally done in the lab by dropping the ethanoyl chloride into concentrated ammonia solution. Don't worry too much about this. Most sources opt out of the problem altogether by quoting the final step + simply as loss of hydrogen ion: " - H ", without making any attempt to explain what removes it.

NUCLEOPHILIC ADDITION / ELIMINATION IN THE REACTION BETWEEN ACYL CHLORIDES AND AMINES
This page gives you the facts and a simple, uncluttered mechanism for the nucleophilic addition / elimination reaction between acyl chlorides (acid chlorides) and amines. If you want the mechanism explained to you in detail, there is a link at the bottom of the page. Ethanoyl chloride is taken as a typical acyl chloride. Any other acyl chloride will behave in the same way. Simply replace the CH3 group in what follows by anything else you want.

Similarly, ethylamine is taken as a typical amine. Any other amine will behave in the same way. Replacing the CH3CH2 group by any other hydrocarbon group won't affect the mechanism in any way.

The reaction between ethanoyl chloride and ethylamine


The facts Ethanoyl chloride reacts violently with a cold concentrated solution of ethylamine. A white solid product is formed which is a mixture of N-ethylethanamide (an N-substituted amide) and ethylammonium chloride.

Notice that, unlike the reactions between ethanoyl chloride and water or ethanol, hydrogen chloride isn't produced - at least, not in any quantity. Any hydrogen chloride formed would immediately react with excess ethylamine to give ethylammonium chloride.

The mechanism The first stage (the addition stage of the reaction) involves a nucleophilic attack on the fairly positive carbon atom by the lone pair on the nitrogen atom in the ethylamine.

The second stage (the elimination stage) happens in two steps. In the first, the carbonoxygen double bond reforms and a chloride ion is pushed off.

That is followed by removal of a hydrogen ion from the nitrogen. This might happen in one of two ways: It might be removed by a chloride ion, producing HCl (which would immediately react with excess ethylamine to give ethylammonium chloride as above) . . .

and

. . . or it might be removed directly by an ethylamine molecule.

The ethylammonium ion, together with the chloride ion already there, makes up the ethylammonium chloride formed in the reaction.

EXPLAINING NUCLEOPHILIC ADDITION / ELIMINATION IN THE REACTION BETWEEN ACYL CHLORIDES AND

AMINES
This page guides you through the mechanism for the nucleophilic addition / elimination reaction between acyl chlorides (acid chlorides) and amines. Ethanoyl chloride is taken as a typical acyl chloride. Similarly, ethylamine is taken as a typical amine. Changing either the acyl choride or the amine won't affect the mechanism in any way.
Important! If you haven't already done so, it would help if you first read the page What is nucleophilic addition / elimination? before you go on. This mechanism also looks extremely complicated! In fact, it isn't any more difficult than the mechanism for the reaction between acyl chlorides and ammonia. Make sure you understand that mechanism before you go any further with this one.

The reaction between ethanoyl chloride and ethylamine


The products of the reaction Remember that the reaction produces N-ethylethanamide and ethylammonium chloride.

N-ethylethanamide is an N-substituted amide. Ethanamide (a simple amide) has the formula CH3CONH2. In an N-substituted amide, one of the hydrogens on the nitrogen has been substituted by a hydrocarbon group - which may be either an alkyl group (as here), or a benzene ring.

In the case we are discussing, an ethyl group has replaced one of the hydrogens on the nitrogen - hence N-ethylethanamide. The other product is ethylammonium chloride. Ammonium chloride is NH4Cl. In ethylammonium chloride, one of the hydrogens on the nitrogen has been substituted by an ethyl group. Ethylamine as a nucleophile A nucleophile is a species (an ion or a molecule) which is strongly attracted to a region of positive charge in something else. Nucleophiles are either fully negative ions, or else have a strongly - charge somewhere on a molecule. Nitrogen is more electronegative than both hydrogen and carbon and so drags the bonding electrons towards itself. That produces a significant amount of negative charge on the nitrogen atom. The nitrogen also has an active lone pair of electrons. It is this which attacks the ethanoyl chloride. Why are acyl chlorides attacked by nucleophiles? The carbon atom in the -COCl group has both an oxygen atom and a chlorine atom attached to it. Both of these are very electronegative. They both pull electrons towards themselves, leaving the carbon atom quite positively charged. It is that carbon atom which is attacked by the lone pair on the nitrogen atom in an ethylamine molecule. The ethanoyl chloride molecule is also planar (flat) around that carbon atom, and that leaves plenty of room for a nucleophile to attack either from above or below the plane of the molecule.

The mechanism The reaction happens in two main stages - an addition stage, followed by an elimination stage. In the addition stage, an ethylamine molecule becomes attached to the carbon in the ethanoyl chloride. As the lone pair on the nitrogen approaches the fairly positive carbon in the ethanoyl chloride, it moves to form a bond with it. In the process, the two electrons in one of the carbon-oxygen bonds are repelled entirely onto the oxygen, leaving it negatively charged.

Notice that the nitrogen atom has gained a positive charge. The underlying reason for this is that when the lone pair forms a bond with the carbon, electrons are moving away from the nitrogen. What matters, though, is that you remember to show the positive charge in an exam. Think of it like this: If you leave out the positive charge, two things are wrong with the equation:

The charges don't balance. You start with two overall neutral molecules and, if you forgot the positive charge, you would end up with a negative ion. There has got to be a positive charge somewhere to balance the negative one. The nitrogen looks wrong! Nitrogen normally forms three bonds, but here it is forming four. Nitrogen can only form four bonds if it carries a positive charge. (A positively charged nitrogen atom has the same electronic structure as carbon - which normally forms four bonds.)

You can put both things right with a positive charge on the nitrogen.

The elimination stage stage of the reaction happens in two steps. In the first, the carbon-oxygen double bond reforms. As the electron pair moves back to form a bond with the carbon, the pair of electrons in the carbon-chlorine bond are forced entirely onto the chlorine to give a chloride ion.

Note: There's nothing sophisticated going on in the switch of the ethyl group from left to right in the diagram. It just makes it look better in the final stages!

Finally, one of the hydrogens attached to the nitrogen is removed as a hydrogen ion. The electrons in the hydrogennitrogen bond move back onto the nitrogen, cancelling the positive charge. The hydrogen ion might be removed in one of two ways. The first way is entirely consistent with what happens in the reactions between water or ethanol and acyl chlorides. The hydrogen is removed by the chloride ion.

The hydrogen chloride produced would at once react with any excess ethylamine present to form ethylammonium chloride.

The other (more likely) possibility is that the hydrogen ion gets removed directly by an ethylamine molecule. This forms an ethylammonium ion. The ethylammonium ion together with the

chloride ion formed in the previous stage makes up the ethylammonium chloride produced in the reaction.

Note: You probably don't really need to know both of these routes for the removal of the hydrogen ion. Choose for yourself. It doesn't matter very much, because both will happen. The route involving the formation of hydrogen chloride is exactly in line with the water or alcohol mechanisms, but you do need to remember to react the HCl with excess ethylamine. You may well feel that this is less daunting than the other route. The removal by another ethylamine molecule is probably the major route as long as the ethylamine is in excess - which it almost certainly will be. The reaction is normally done in the lab by dropping the ethanoyl chloride into concentrated ethylamine solution. Don't worry too much about this. Most sources opt out of the problem altogether by quoting the final step simply as loss of + hydrogen ion: " - H ", without making any attempt to explain what removes it.

Mechanisms described elsewhere on the site


These are discussed in a section on acid catalysis in organic chemistry. The hydration of ethene to make ethanol . . . The esterification reaction . . . The acid catalysed hydrolysis of esters . . .

THE MECHANISM FOR THE ACID CATALYSED HYDRATION OF ETHENE


This page describes the mechanism for the hydration of ethene to make ethanol using phosphoric(V) acid as the catalyst.

The hydration of ethene to make ethanol


A reminder of the facts Ethene is mixed with steam and passed over a catalyst consisting of solid silicon dioxide coated with phosphoric(V) acid. The temperature used is 300C and the pressure is about 60 to 70 atmospheres.

Note: If you are interested in the reasons for the conditions used in this reaction, you will find them in the equilibrium section of this site by following this link. Use the BACK button on your browser to return to this page.

The mechanism for the hydration of ethene This assumes that you know about the electrophilic addition reactions of ethene, and about the use of curly arrows in organic reaction mechanisms. If you aren't happy about either of these follow the link below before you go any further.
Note: Follow this link for the electrophilic addition reactions of ethene. You will find a link explaining the use of curly arrows in mechanisms on that page if you need it. These pages are in a completely different part of the site. The quickest way to return here is to use the BACK button on your browser, or the history file or the Go menu if you get seriously

waylaid!

All the steps in the mechanism below are shown as one-way reactions because it makes the mechanism look less confusing. It doesn't affect the argument, but in fact all the steps are reversible. Step 1 All of the hydrogen atoms in the phosphoric(V) acid are fairly positively charged because they are attached to a very electronegative oxygen atom.
Note: If you aren't sure about electronegativity, you might like to follow this link. Use the BACK button on your browser to return quickly to this page.

One of these hydrogens is strongly attracted to the carboncarbon double bond. The pi part of the bond breaks and the electrons in it move down to make a new bond with the hydrogen atom. That forces the electrons in the hydrogenoxygen bond down entirely onto the oxygen.

Note: It is easy to see why the oxygen carries a negative charge. It has gained full control over the electron pair in the original bond - so has acquired an extra electron which originally belonged to the hydrogen. The carbon atom has a positive charge because one of the electrons in the pi bond originally belonged to it. It loses that electron when the

pi bond breaks.

Step 2 The carbocation (carbonium ion) formed reacts with one of the lone pairs on a water molecule. A carbocation is one which carries a positive charge on a carbon atom.

Note: The easiest way of remembering that the oxygen has to carry a positive charge is that you are reacting a positive ion with a neutral molecule. That means that there must be a positive charge on the product somewhere. The only way that an oxygen atom can be joined to three things at the same time is if it carries a positive charge. A positive charge gives the oxygen the same electronic structure as a nitrogen atom which can, of course, form three bonds.

Step 3 Finally, one of the hydrogens on the oxygen is removed by reaction with the dihydrogenphosphate(V) ion, H2PO4-, formed in the first step.

The phosphoric(V) acid catalyst has been regenerated.


Note: This is a bit of a simplification. The concentrated

phosphoric(V) acid probably contains some water, in addition to the water in the steam. It is likely that some of the hydrogen ions in this last step will be removed by water molecules, and possibly transferred to a dihydrogenphosphate(V) ion to make phosphoric(V) acid in a later step. Other hydrogen ions will be removed directly by the dihydrogenphosphate(V) ions as shown above. Don't worry about this. Your examiners will almost certainly not be looking for this sort of fine detail.

The reversibility of this reaction As mentioned before, this reaction is entirely reversible - and so each step is reversible. In the reverse direction, this is a dehydration of ethanol to make ethene. If you are interested, you will find the mechanism written in the reverse direction on another page.
Note: This link will lead you to the reverse of this hydration process, the dehydration of ethanol. The catalyst used on that page is concentrated sulphuric acid rather than phosphoric(V) acid, because that is the acid often used in the lab. You could write either of these mechanisms (in either direction) perfectly well using the structure of either of these acids.

THE MECHANISM FOR THE ESTERIFICATION REACTION


This page looks in detail at the mechanism for the formation of esters from carboxylic acids and alcohols in the presence of concentrated sulphuric acid acting as the catalyst. It uses the formation of ethyl ethanoate from ethanoic acid and ethanol as a typical example.

The mechanism for the formation of ethyl ethanoate


A reminder of the facts Ethanoic acid reacts with ethanol in the presence of concentrated sulphuric acid as a catalyst to produce the ester, ethyl ethanoate. The reaction is slow and reversible. To reduce

the chances of the reverse reaction happening, the ester is distilled off as soon as it is formed.

The mechanism
Warning! This is a fairly complex mechanism, and is definitley NOT required for any UK A level (or equivalent) syllabus. I have included it in case it is of use to my many non-UK visitors.

All the steps in the mechanism below are shown as one-way reactions because it makes the mechanism look less confusing. The reverse reaction is actually done sufficiently differently that it affects the way the mechanism is written. You will find a link to the hydrolysis of esters further down the page if you are interested.
Note: The explanation assumes that you know about the use of curly arrows in organic reaction mechanisms. If you aren't happy about these follow this link before you go any further. (To be honest, if you are that unsure about the conventions used in reaction mechanisms, you probably shouldn't be reading this page anyway - you will find it distinctly scary!) Use the BACK button on your browser to return to this page.

Step 1 In the first step, the ethanoic acid takes a proton (a hydrogen ion) from the concentrated sulphuric acid. The proton becomes attached to one of the lone pairs on the oxygen which is doublebonded to the carbon.

The transfer of the proton to the oxygen gives it a positive charge, but it is actually misleading to draw the structure in this way (although nearly everybody does!). The positive charge is delocalised over the whole of the righthand end of the ion, with a fair amount of positiveness on the carbon atom. In other words, you can think of an electron pair shifting to give this structure:

You could also imagine another electron pair shift producing a third structure:

So which of these is the correct structure of the ion formed? None of them! The truth lies somewhere in between all of them. One way of writing the delocalised structure of the ion is like this:

The double headed arrows are telling you that each of the individual structures makes a contribution to the real structure of the ion. They don't mean that the bonds are flipping back and forth between one structure and another. The various structures are known as resonance structures or canonical forms. There will be some degree of positive charge on both of the oxygen atoms, and also on the carbon atom. Each of the bonds between the carbon and the two oxygens will be the same somewhere between a single bond and a double bond.
Note: You will find a more pictorial look at a similar case to this in a page discussing the acidity of organic acids. Amongst other things, that page

looks at the structure of ions like the ethanoate ions which also have delocalised charges. Use the BACK button on your browser to return easily to this page.

For the purposes of the rest of this discussion, we are going to use the structure where the positive charge is on the carbon atom. Step 2 The positive charge on the carbon atom is attacked by one of the lone pairs on the oxygen of the ethanol molecule.

Note: You could work out precisely why that particular oxygen carries the positive charge on the right-hand side. On the other hand, you could realise that there has to be a positive charge somewhere (because you started with one), and that particular oxygen doesn't look right - it has too many bonds. Put the charge on there! That's a quick rough-and-ready reasoning which works every time I use it!

Step 3 What happens next is that a proton (a hydrogen ion) gets transferred from the bottom oxygen atom to one of the others. It gets picked off by one of the other substances in the mixture (for example, by attaching to a lone pair on an unreacted ethanol molecule), and then dumped back onto one of the oxygens more or less at random. The net effect is:

Step 4 Now a molecule of water is lost from the ion.

The product ion has been drawn in a shape to reflect the product which we are finally getting quite close to! The structure for the latest ion is just like the one we discusssed at length back in step 1. The positive charge is actually delocalised all over that end of the ion, and there will also be contributions from structures where the charge is on the either of the oxygens:

It is easier to follow what is happening if we keep going with the structure with the charge on the carbon. Step 5 The hydrogen is removed from the oxygen by reaction with the hydrogensulphate ion which was formed way back in the first step.

And there we are! The ester has been formed, and the sulphuric acid catalyst has been regenerated.

THE MECHANISM FOR THE ACID CATALYSED HYDROLYSIS OF ESTERS

This page looks in detail at the mechanism for the hydrolysis of esters in the presence of a dilute acid (such as hydrochloric acid or sulphuric acid) acting as the catalyst. It uses ethyl ethanoate as a typical ester.

The mechanism for the hydrolysis of ethyl ethanoate A reminder of the facts Ethyl ethanoate is heated under reflux with a dilute acid such as dilute hydrochloric acid or dilute sulphuric acid. The ester reacts with the water present to produce ethanoic acid and ethanol. Because the reaction is reversible, an equilibrium mixture is produced containing all four of the substances in the equation. In order to get as much hydrolysis as possible, a large excess of water can be used. The dilute acid provides both the acid catalyst and the water.

The mechanism
Warning! This is a fairly complex mechanism, and is definitley NOT required for any UK A level (or equivalent) syllabus. I have included it in case it is of use to my many non-UK visitors.

All the steps in the mechanism below are shown as one-way

reactions because it makes the mechanism look less confusing. The reverse reaction is actually done sufficiently differently that it affects the way the mechanism is written. You will find a link to the esterification reaction further down the page if you are interested.
Note: The explanation assumes that you know about the use of curly arrows in organic reaction mechanisms. If you aren't happy about these follow this link before you go any further. (To be honest, if you are that unsure about the conventions used in reaction mechanisms, you probably shouldn't be reading this page anyway - you will find it distinctly scary!) Use the BACK button on your browser to return to this page.

Step 1 The actual catalyst in this case is the hydroxonium ion, H3O+, present in all solutions of acids in water. In the first step, the ester takes a proton (a hydrogen ion) from the hydroxonium ion. The proton becomes attached to one of the lone pairs on the oxygen which is double-bonded to the carbon.

The transfer of the proton to the oxygen gives it a positive charge, but the charge is actually delocalised (spread around) much more widely than this shows. One way of representing this delocalisation is to draw a number of structures called resonance structures or canonical forms joined by double-headed arrows. You could, if you wished, put in some curly arrows to show the movements of electrons which change one of these structures into the next.

None of these formulae represents the true structure of the ion but each gives you some information about it. For example, notice where the positive charge is in the three structures. What this means in reality is that the positive charge is spread around over those three atoms - the two oxygens and the carbon.
Note: If you haven't come across canonical forms as a way of representing delocalisation, don't worry about it particularly. It is important, however, that you don't imagine that the molecule is rapidly flipping from one structure to another. The double-headed arrows mean something different. A mule is a hybrid of a donkey and a horse. In this notation, you could represent a mule by writing donkey and horse connected by a doubleheaded arrow. Neitherdonkey nor horse accurately represents what a mule looks like, but with a bit of imagination you could build up a fairly good picture of a mule by combining together the characteristics of both donkey and horse. But a mule obviously doesn't spend its time rapidly changing back and forth between being a donkey and a horse!

The next stage of the mechanism involves an attack on the carbon, and so it is convenient to use the structure showing the positive charge on that carbon in the next step. Step 2 The positive charge on the carbon atom is attacked by one of the lone pairs on the oxygen of a water molecule.

Note: You could work out precisely why that particular oxygen carries the positive charge on the right-hand side. On the other hand, you could realise that there has to be a positive charge somewhere (because you started with one), and that particular oxygen doesn't look right - it has too many bonds. Put the charge on there! That's a quick rough-and-ready reasoning which works every time I use it!

Step 3 What happens next is that a proton (a hydrogen ion) gets transferred from the bottom oxygen atom to one of the others. It gets picked off by one of the other substances in the mixture (for example, by attaching to a lone pair on a water molecule), and then dumped back onto one of the oxygens more or less at random. Eventually, by chance, it will join to the oxygen with the ethyl group attached. When that happens, the net effect is:

Step 4 Now a molecule of ethanol is lost from the ion. That's one of the products of the reaction.

The structure for the latest ion is just like the one we discusssed at length back in step 1. The positive charge is actually delocalised all over that end of the ion. The real structure will be

a hybrid of these:

It is easier to follow what is happening if we keep going with the structure with the charge on the carbon. Step 5 The hydrogen is removed from the oxygen by reaction with a water molecule.

And there we are! We have produced the ethanoic acid (the other product of the reaction) and the hydroxonium ion catalyst has been regenerated.

Potrebbero piacerti anche