You are on page 1of 40

Chemistry Unit 5

Rates
Collision theory In order to react in the first place, minus all the temperature and surface area malarkey, to particles must collide with each other and they may react. May react? Isnt it enough to have two particles colliding with each other? Well no they have to collide the correct way round AND with the right energy for the bonds to break. Orientation o Consider a simple reaction between ethene CH2=CH2 and HCL they react to form chloroethane. o As a result of the collision a double bond is broken and a single bond if formed o The reaction can only happen if the end of the hydrogen approaches the carbon-carbon double bond, any other orientation and the reaction cannot happen o In unsymmetrical compounds the orientation is very important as to whether the reaction will take place Activation energy o Even if the particles are the right way round they still wont react unless they have a minimum amount of energy called the activation energy. Surface Area The more finely divided the solid is, the faster the reaction happens. A powdered solid will normally produce a faster reaction than if the same mass is present as a single lump. The powdered solid has a greater surface area than the single lump. The explanation o As we know from the previous section particles has to collide in order to react so if a particle has more surface area there is more places for the particles to collide and there fore react o Consider the reaction between hydrochloric acid and magnesium o Increasing the speed of collisions increases the rate of reaction

Concentration Simply if you increase the concentration of the reactants you increase the rate of reaction. Pressure Increasing the pressure on a reaction involving reacting gases increases the rate of reaction. Changing the pressure on a reaction, which involves only solids or liquids, has no effect on the rate. Temperature Increasing temperature increases the rate of reaction, because when you increase the temperature of a substance this increases the kinetic energy of the particles within it. (They move faster) this increases the amount of collisions and more collisions have the correct activation energy, so there fore an increase in the rate of reaction. Catalysts Catalysts speed up rate of reaction by creating an alternate reaction pathway with lower activation energy. Usually by creating intermediate compounds, the most important thing about catalysts is that they arent used up in the reaction or if they are they are recreated. Catalysts can be homogeneous(same state as reactants) or heterogeneous ( different state)

Rate equations

Measuring rates COLORIMETRY: Measures change in intensity of colour CLOCK REACTIONS: Time taken for a particular amount of reactant to react//product to be formed MASS CHANGE/VOLUME CHANGE: Used for gas reactants/products TITRIMETRIC ANALYSIS: Measures changes in conc of reactant/product when one is acid, alkali or iodine by reacting with a known volume of standard solution to neutralise Calculating rate equations Average rate = the change in concentration over time Rate = k[A]x[B]y When asked in a question to work out the rate equation, you will be given a table of values, when the concentration of substance A remains constant, you calculate how much B increases by (twice, 3x, 4x) and how much the rate in creases by. If the rate increase the same amount this is first order.( if the rate increases the square (2 and 4, 3 and 9) of the concentration of B this is second order) K = rate constant, constant at a particular temperature

[A] and [B] = concentrations of substances A and B x and y = partial orders Overall order = x + y The Arrhenius equation

luckily we dont need to memorise this but its nice to understand it so you dont panic in the exam so here are what the various symbols mean Starting with the easy ones . . . Temperature, o T to fit into the equation, this has to be measured in kelvin. The gas constant, R o This is a constant which comes from an equation, pV=nRT, which relates the pressure, volume and temperature of a particular number of moles of gas. It turns up in all sorts of unlikely places! Activation energy, EA o This is the minimum energy needed for the reaction to occur. To fit this into the equation, it has to be expressed in joules per mole - not in kJ mol-1 You may also come across it in a different form created by a mathematical operation on the standard one:

"ln" is a form of logarithm. Don't worry about what it means. If you need to use this equation, just find the "ln" button on your calculator.

Entropy
A measure of disorder Zero entropy = perfectly crystalline solid at 0k 2nd law of thermodynamics = entropy always increases Stotal = Ssystem + Ssurroundings Entropy of System Ssystem = Sproducts - Sreactants Use of a data booklet Entropy of surroundings Ssurroundings= -H/T Units =JK-1mol-1 Factors affecting entropy Physical state: solid < liquid < gas o Particles get further apart, so there are more possibilities for their position in space Dissolving: o Entropy increases if a solid is dissolved, entropy can either increase or decrease if a liquid is dissolved Number of particles: o More particles means higher entropy because there are more the particles arrangement in space Temperature: o Increased temperature increases entropy because it increases the quanta of energy in particles, so there are more distributions of the particles

Change of state: o Rapid change in entropy because particles move further apart/ closer together Feasibility Total entropy: o Must increase and must be positive for a reaction to be possible Increasing Temperature: o increases entropy, so can be used to make a reaction feasible Dissolving Enthalpy of solution: o Enthalpy change when 1 mole of solute is dissolved in sufficient solvent to produce an infinitely dilute solution Lattice enthalpy: o Enthalpy change when 1 mol of a solid ionic compound is produced from gaseous ions that start infinitely far apart Enthalpy of hydration: o Enthalpy change when 1 mol of aqueous ions is formed from gaseous ions Enthalpy of solution: o Sum of enthalpy of solutions Factors affecting lattice enthalpy Charge of ions: o Larger charge on ions increases lattice enthalpy Ionic radii: o Larger ionic radii decreases lattice enthalpy

Equilibrium

Explaining the basics a reversible reaction is a chemical change in which the products can be converted back to the original reactants under suitable conditions. Shown by the sign When a reversible reaction occurs in a closed system, a chemical equilibrium is formed o A closed system means nothing can enter or leave o Eventually the system settles down and a state of concentration balance exsists o BUT! The reactions dont stop! Reactants are continually forming products, and products are re-forming reactants, hence dynamic equilibrium. Le chateliers Principle If changing the conditions disturbs a dynamic equilibrium, the position of equilibrium moves to counteract the change. Rule 1 raising the temperature favours the endothermic direction and lowering favours the exothermic Rule 2 increasing the pressure favours the side of the equation with the least number or gaseous molecules Rule 3 a catalyst doesnt affect the position of an equilibrium BUT it does help equilibrium be reached faster Writing an expression for Kc We are going to look at a general case with the equation: No state symbols have been given, but they will be all (g), or all (l), or all (aq) if the reaction was between substances in solution in water. If you allow this reaction to reach equilibrium and then measure the

equilibrium concentrations of everything, you can combine these concentrations into an expression known as equilibrium constant. The equilibrium constant always has the same value (provided you don't change the temperature), irrespective of the amounts of A, B, C and D you started with. It is also unaffected by a change in pressure or whether or not you are using a catalyst.

Compare this with the chemical equation for the equilibrium. The convention is that the substances on the right-hand side of the equation are written at the top of the Kc expression, and those on the left-hand side at the bottom. The indices (the powers that you have to raise the concentrations to - for example, squared or cubed or whatever) are just the numbers that appear in the equation Kc is governed by temperature, this is only factor that can change K c o If the forward reaction is exothermic, Kc will decrease in value with temperature increase o If the forward reaction is endothermic Kc will increase in value with increase in temperature If Kc is > 105 the posision of equilibrium is on the right hand side, almost 100% completion If Kc is aproximetly 1 the equilibrium is dead centre (dynamic equilibrium) If Kc is < 10-5 (Very small) the equilibrium lies on the left hand side, very low RHS yield, backwards reaction is favoured Kp expression Mole fraction If you have a mixture of gases (A, B, C, etc), then the mole fraction of gas A is worked out by dividing the number of moles of A by the total number of moles of gas. The mole fraction of gas A is often given the symbol xA. The mole fraction of gas B would be xB - and so on

To calculate the partial pressure of a gas in order to use in K p you times

the total pressure by the mole fraction.

ACID-BASE Equilibrium

The Arrhenius theory The theory Acids are substances which produce hydrogen ions in solution. Bases are substances which produce hydroxide ions in solution. Neutralisation happens because hydrogen ions and hydroxide ions react to produce water Limitations of the theory Hydrochloric acid is neutralised by both sodium hydroxide solution and ammonia solution. In both cases, you get a colourless solution which you can crystallise to get a white salt - either sodium chloride or ammonium chloride. These are clearly very similar reactions. The full equations are: In the sodium hydroxide case, hydrogen ions from the acid are reacting with hydroxide ions from the sodium hydroxide - in line with the Arrhenius theory. However, in the ammonia case, there don't appear to be any hydroxide ions! You can get around this by saying that the ammonia reacts with the water it is dissolved in to produce ammonium ions and hydroxide ions: This is a reversible reaction, and in a typical dilute ammonia solution, about 99% of the ammonia remains as ammonia molecules. Nevertheless,

there are hydroxide ions there, and we can squeeze this into the Arrhenius theory. However, this same reaction also happens between ammonia gas and hydrogen chloride gas. In this case, there aren't any hydrogen ions or hydroxide ions in solution - because there isn't any solution. The Arrhenius theory wouldn't count this as an acid-base reaction, despite the fact that it is producing the same product as when the two substances were in solution. That's silly! The bronsted-lowry acid theory The theory An acid is a proton (hydrogen ion) donor. A base is a proton (hydrogen ion) acceptor. The relationship between the Bronsted-Lowry theory and the Arrhenius theory The Bronsted-Lowry theory doesn't go against the Arrhenius theory in any way - it just adds to it. Hydroxide ions are still bases because they accept hydrogen ions from acids and form water. An acid produces hydrogen ions in solution because it reacts with the water molecules by giving a proton to them. When hydrogen chloride gas dissolves in water to produce hydrochloric acid, the hydrogen chloride molecule gives a proton (a hydrogen ion) to a water molecule. A co-ordinate (dative covalent) bond is formed between one of the lone pairs on the oxygen and the hydrogen from the HCl. Hydroxonium ions, H3O+, are produced.

When an acid in solution reacts with a base, what is actually functioning as the acid is the hydroxonium ion. For example, a proton is transferred from a hydroxonium ion to a hydroxide ion to make water. Showing the electrons, but leaving out the inner ones:

It is important to realise that whenever you talk about hydrogen ions in

solution, H+(aq), what you are actually talking about are hydroxonium ions. Conjugate pairs When hydrogen chloride dissolves in water, almost 100% of it reacts with the water to produce hydroxonium ions and chloride ions. Hydrogen chloride is a strong acid, and we tend to write this as a one-way reaction: n fact, the reaction between HCl and water is reversible, but only to a very minor extent. In order to generalise, consider an acid HA, and think of the reaction as being reversible. Thinking about the forward reaction: The HA is an acid because it is donating a proton (hydrogen ion) to the water. The water is a base because it is accepting a proton from the HA. But there is also a back reaction between the hydroxonium ion and the Aion: The H3O+ is an acid because it is donating a proton (hydrogen ion) to the A- ion. The A- ion is a base because it is accepting a proton from the H 3O+. The reversible reaction contains two acids and two bases. We think of them in pairs, called conjugate pairs.

When the acid, HA, loses a proton it forms a base, A -. When the base, A-, accepts a proton back again, it obviously refoms the acid, HA. These two are a conjugate pair. Members of a conjugate pair differ from each other by the presence or absence of the transferable hydrogen ion. If you are thinking about HA as the acid, then A- is its conjugate base. If you are thinking about A- as the base, then HA is its conjugate acid. The water and the hydroxonium ion are also a conjugate pair. Thinking of the water as a base, the hydroxonium ion is its conjugate acid because it has the extra hydrogen ion which it can give away again. Thinking about the hydroxonium ion as an acid, then water is its conjugate base. The water can accept a hydrogen ion back again to reform the hydroxonium ion. A second example of conjugate pairs This is the reaction between ammonia and water that we looked at earlier:

Think first about the forward reaction. Ammonia is a base because it is accepting hydrogen ions from the water. The ammonium ion is its conjugate acid - it can release that hydrogen ion again to reform the ammonia. The water is acting as an acid, and its conjugate base is the hydroxide ion. The hydroxide ion can accept a hydrogen ion to reform the water. Looking at it from the other side, the ammonium ion is an acid, and ammonia is its conjugate base. The hydroxide ion is a base and water is its conjugate acid. Amphoteric substances You may possibly have noticed (although probably not!) that in one of the last two examples, water was acting as a base, whereas in the other one it was acting as an acid. A substance which can act as either an acid or a base is described as being amphoteric.

Strong Acids Explaining the term strong acid When an acid dissolves in water, a proton (hydrogen ion) is transferred to a water molecule to produce a hydroxonium ion and a negative ion depending on what acid you are starting from. In the general case . . . These reactions are all reversible, but in some cases, the acid is so good at giving away hydrogen ions that we can think of the reaction as being one-way. The acid is virtually 100% ionised. For example, when hydrogen chloride dissolves in water to make hydrochloric acid, so little of the reverse reaction happens that we can write: At any one time, virtually 100% of the hydrogen chloride will have reacted to produce hydroxonium ions and chloride ions. Hydrogen chloride is described as a strong acid. A strong acid is one which is virtually 100% ionised in solution. Other common strong acids include sulphuric acid and nitric acid. You may find the equation for the ionisation written in a simplified form:

This shows the hydrogen chloride dissolved in the water splitting to give hydrogen ions in solution and chloride ions in solution. This version is often used in this work just to make things look easier. If you use it, remember that the water is actually involved, and that when you write H+(aq) what you really mean is a hydroxonium ion, H3O+. Strong acids and pH pH is a measure of the concentration of hydrogen ions in a solution. Strong acids like hydrochloric acid at the sort of concentrations you normally use in the lab have a pH around 0 to 1. The lower the pH, the higher the concentration of hydrogen ions in the solution. Defining pH

Working out the pH of a strong acid Suppose you had to work out the pH of 0.1 mol dm -3 hydrochloric acid. All you have to do is work out the concentration of the hydrogen ions in the solution, and then use your calculator to convert it to a pH. With strong acids this is easy. Hydrochloric acid is a strong acid - virtually 100% ionised. Each mole of HCl reacts with the water to give 1 mole of hydrogen ions and 1 mole of chloride ions That means that if the concentration of the acid is 0.1 mol dm -3, then the concentration of hydrogen ions is also 0.1 mol dm -3. Use your calculator to convert this into pH. My calculator wants me to enter 0.1, and then press the "log" button. Yours might want you to do it in a different order. You need to find out! log10 [0.1] = -1 But pH = - log10 [0.1] - (-1) = 1 The pH of this acid is 1 Weak acids Explaining the term "weak acid" A weak acid is one which doesn't ionise fully when it is dissolved in water. Ethanoic acid is a typical weak acid. It reacts with water to produce hydroxonium ions and ethanoate ions, but the back reaction is more successful than the forward one. The ions react very easily to reform the acid and the water. At any one time, only about 1% of the ethanoic acid molecules have

converted into ions. The rest remain as simple ethanoic acid molecules. Most organic acids are weak. Hydrogen fluoride (dissolving in water to produce hydrofluoric acid) is a weak inorganic acid that you may come across elsewhere. Comparing the strengths of weak acids The position of equilibrium of the reaction between the acid and water varies from one weak acid to another. The further to the left it lies, the weaker the acid is. The acid dissociation constant, Ka You can get a measure of the position of an equilibrium by writing an equilibrium constant for the reaction. The lower the value for the constant, the more the equilibrium lies to the left. The dissociation (ionisation) of an acid is an example of a homogeneous reaction. Everything is present in the same phase - in this case, in solution in water. You can therefore write a simple expression for the equilibrium constant, Kc. Here is the equilibrium again: You might expect the equilibrium constant to be written as:

However, if you think about this carefully, there is something odd about it. At the bottom of the expression, you have a term for the concentration of the water in the solution. That's not a problem - except that the number is going to be very large compared with all the other numbers. In 1 dm3 of solution, there are going to be about 55 moles of water. If you had a weak acid with a concentration of about 1 mol dm -3, and only about 1% of it reacted with the water, the number of moles of water is only going to fall by about 0.01. In other words, if the acid is weak the concentration of the water is virtually constant. In that case, there isn't a lot of point in including it in the expression as if it were a variable. Instead, a new equilibrium constant is defined which leaves it out. This new equilibrium constant is called Ka.

now you can use this to calculate the pH for a weak acid, we can assume [H3O+]=[A-] and we already know that the hydroxonium ion is just H+ in water, so if we are given Ka we can rearrange this equation to calculate the H+ concentration which we can then place into the log equation above for pH.

An introduction to pKa pKa bears exactly the same relationship to Ka as pH does to the hydrogen ion concentration: Remember this: The lower the value for pKa, the stronger the acid. The higher the value for pKa, the weaker the acid. The ionic product of water, KW and PKw The important equilibrium in water Water molecules can function as both acids and bases. One water molecule (acting as a base) can accept a hydrogen ion from a second one (acting as an acid). This will be happening anywhere there is even a trace of water - it doesn't have to be pure. A hydroxonium ion and a hydroxide ion are formed.

However, the hydroxonium ion is a very strong acid, and the hydroxide ion is a very strong base. As fast as they are formed, they react to produce water again. The net effect is that an equilibrium is set up. At any one time, there are incredibly small numbers of hydroxonium ions and hydroxide ions present. Further down this page, we shall calculate the concentration of hydroxonium ions present in pure water. It turns out to be 1.00 x 10-7 mol dm-3 at room temperature. You may well find this equilibrium written in a simplified form: This is OK provided you remember that H+(aq) actually refers to a hydroxonium ion. Defining the ionic product for water, Kw Kw is essentially just an equilibrium constant for the reactions shown. You may meet it in two forms: Based on the fully written equilibrium . . . . . . or on the simplified equilibrium: You may find them written with or without the state symbols. Whatever version you come across, they all mean exactly the same thing!

You may wonder why the water isn't written on the bottom of these equilibrium constant expressions. So little of the water is ionised at any one time, that its concentration remains virtually unchanged - a constant. Kw is defined to avoid making the expression unnecessarily complicated by including another constant in it The value of Kw Like any other equilibrium constant, the value of Kw varies with temperature. Its value is usually taken to be 1.00 x 10 -14 mol2 dm-6 at room temperature. In fact, this is its value at a bit less than 25C. pKw The relationship between Kw and pKw is exactly the same as that between Ka and pKa, or [H+] and pH. The Kw value of 1.00 x 10-14 mol2 dm-6 at room temperature gives you a pKw value of 14. Try it on your calculator! Notice that pKw doesn't have any units Strong Bases Explaining the term "strong base" A strong base is something like sodium hydroxide or potassium hydroxide which is fully ionic. You can think of the compound as being 100% split up into metal ions and hydroxide ions in solution. Each mole of sodium hydroxide dissolves to give a mole of hydroxide ions in solution.

Some strong bases like calcium hydroxide aren't very soluble in water. That doesn't matter - what does dissolve is still 100% ionised into calcium ions and hydroxide ions. Calcium hydroxide still counts as a strong base because of that 100% ionisation. Working out the pH of a strong base Remember that: Since pH is a measure of hydrogen ion concentration, how can a solution which contains hydroxide ions have a pH? To understand this, you need to know about the ionic product for water. An outline of the method of working out the pH of a strong base Work out the concentration of the hydroxide ions. Use Kw to work out the hydrogen ion concentration. Convert the hydrogen ion concentration to a pH. Weak bases

Explaining the term "weak base" Ammonia is a typical weak base. Ammonia itself obviously doesn't contain hydroxide ions, but it reacts with water to produce ammonium ions and hydroxide ions. However, the reaction is reversible, and at any one time about 99% of the ammonia is still present as ammonia molecules. Only about 1% has actually produced hydroxide ions. A weak base is one which doesn't convert fully into hydroxide ions in solution. Comparing the strengths of weak bases in solution: Kb When a weak base reacts with water, the position of equilibrium varies from base to base. The further to the left it is, the weaker the base.

You can get a measure of the position of equilibrium by writing an equilibrium constant for the reaction. The lower the value for the constant, the more the equilibrium lies to the left. In this case the equilibrium constant is called Kb. This is defined as:

again just like with weak acids you can use this to calculate pH, again the top is equal so you can rearrange to calculate OH- concentration then using Kw and the log equation you can calculate pH pKb The relationship between Kb and pKb is exactly the same as all the other "p" terms in this topic: As Kb gets bigger, pKb gets smaller. The lower the value of pKb, the stronger the base. This is exactly in line with the corresponding term for acids, pK a - the smaller the value, the stronger the acid.

pH Curves
Equivalence point Sorting out some confusing terms When you carry out a simple acid-base titration, you use an indicator to tell you when you have the acid and alkali mixed in exactly the right proportions to "neutralise" each other. When the indicator changes colour, this is often described as the end point of the titration. In an ideal world, the colour change would happen when you mix the two solutions together in exactly equation proportions. That particular mixture is known as the equivalence point. For example, if you were titrating sodium hydroxide solution with hydrochloric acid, both with a concentration of 1 mol dm-3, 25 cm3 of sodium hydroxide solution would need exactly the same volume of the acid - because they react 1 : 1 according to the equation. In this particular instance, this would also be the neutral point of the titration, because sodium chloride solution has a pH of 7. But that isn't necessarily true of all the salts you might get formed. For example, if you titrate ammonia solution with hydrochloric acid, you would get ammonium chloride formed. The ammonium ion is slightly acidic, and so pure ammonium chloride has a slightly acidic pH. That means that at the equivalence point (where you had mixed the solutions in the correct proportions according to the equation), the solution wouldn't actually be neutral. To use the term "neutral point" in this context would be misleading. Similarly, if you titrate sodium hydroxide solution with ethanoic acid, at the equivalence point the pure sodium ethanoate formed has a slightly alkaline pH because the ethanoate ion is slightly basic. To summarise: The term "neutral point" is best avoided. The term "equivalence point" means that the solutions have been mixed in exactly the right proportions according to the equation. The term "end point" is where the indicator changes colour. As you will see on the page about indicators, that isn't necessarily exactly the same as the equivalence point. Simple pH curves

All the following titration curves are based on both acid and alkali having a concentration of 1 mol dm-3. In each case, you start with 25 cm 3 of one of the solutions in the flask, and the other one in a burette. Although you normally run the acid from a burette into the alkali in a flask, you may need to know about the titration curve for adding it the other way around as well. Alternative versions of the curves have been described in most cases.

Titration curves for strong acid v strong base We'll take hydrochloric acid and sodium hydroxide as typical of a strong acid and a strong base. Running acid into the alkali

You can see that the pH only falls a very small amount until quite near the equivalence point. Then there is a really steep plunge. If you calculate the values, the pH falls all the way from 11.3 when you have added 24.9 cm 3 to 2.7 when you have added 25.1 cm3. Running alkali into the acid This is very similar to the previous curve except, of course, that the pH starts off low and increases as you add more sodium hydroxide solution.

Again, the pH doesn't change very much until you get close to the equivalence point. Then it surges upwards very steeply. Titration curves for strong acid v weak base This time we are going to use hydrochloric acid as the strong acid and ammonia solution as the weak base.

Running acid into the alkali

Because you have got a weak base, the beginning of the curve is obviously going to be different. However, once you have got an excess of acid, the curve is essentially the same as before. At the very beginning of the curve, the pH starts by falling quite quickly as the acid is added, but the curve very soon gets less steep. This is because a buffer solution is being set up - composed of the excess ammonia and the ammonium chloride being formed. Notice that the equivalence point is now somewhat acidic ( a bit less than pH 5), because pure ammonium chloride isn't neutral. However, the equivalence point still falls on the steepest bit of the curve. That will turn out to be important in choosing a suitable indicator for the titration. Running alkali into the acid At the beginning of this titration, you have an excess of hydrochloric acid. The shape of the curve will be the same as when you had an excess of acid at the start of a titration running sodium hydroxide solution into the acid. It is only after the equivalence point that things become different. A buffer solution is formed containing excess ammonia and ammonium chloride. This resists any large increase in pH - not that you would expect a very large increase anyway, because ammonia is only a weak base.

Titration curves for weak acid v strong base We'll take ethanoic acid and sodium hydroxide as typical of a weak acid

and a strong base. Running acid into the alkali For the first part of the graph, you have an excess of sodium hydroxide. The curve will be exactly the same as when you add hydrochloric acid to sodium hydroxide. Once the acid is in excess, there will be a difference.

Past the equivalence point you have a buffer solution containing sodium ethanoate and ethanoic acid. This resists any large fall in pH. Running alkali into the acid

The start of the graph shows a relatively rapid rise in pH but this slows down as a buffer solution containing ethanoic acid and sodium ethanoate is produced. Beyond the equivalence point (when the sodium hydroxide is in excess) the curve is just the same as that end of the HCl - NaOH graph. Titration curves for weak acid v weak base The common example of this would be ethanoic acid and ammonia. It so happens that these two are both about equally weak - in that case, the equivalence point is approximately pH 7. Running acid into the alkali This is really just a combination of graphs you have already seen. Up to the equivalence point it is similar to the ammonia - HCl case. After the equivalence point it is like the end of the ethanoic acid - NaOH curve.

Notice that there isn't any steep bit on this graph. Instead, there is just what is known as a "point of inflexion". That lack of a steep bit means that it is difficult to do a titration of a weak acid against a weak base. Calculations based on these graphs To calculate the equivalence point you (using a ruler) measure the steep bit, divide this value by two, mark a dot at this number, follow across to find pH, down for the volume of acid added. At half equivalence point pH=pKa

Acid-Base indicators

METHYL ORANGE: Use when there is a strong acid PHENOLPHTHALEIN: Use when there is a strong alkali NO INDICATOR AVAILABLE: When both the acid and alkali are weak, as there is no sharp change in pH USE EITHER: When strong acid and strong alkali, as there is a long, sharp change in pH Indicators are weak acids, The un-ionised indicator is colour 1 , whereas the ion is colour 2 Using litmus as example of how this works Adding hydroxide ions:

Adding hydrogen ions:

If the concentrations of HLit and Lit - are equal: At some point during the movement of the position of equilibrium, the concentrations of the two colours will become equal. The colour you see will be a mixture of the two.

Buffers
Definition A buffer solution is one, which resists changes in pH when small quantities of an acid or an alkali are added to it. Acidic buffer solutions An acidic buffer solution is simply one, which has a pH less than 7. Acidic buffer solutions are commonly made from a weak acid and one of its salts - often a sodium salt. A common example would be a mixture of ethanoic acid and sodium ethanoate in solution. In this case, if the solution contained equal molar concentrations of both the acid and the salt, it would have a pH of 4.76. It wouldn't matter what the concentrations were, as long as they were the same. You can change the pH of the buffer solution by changing the ratio of acid to salt, or by choosing a different acid and one of its salts. Alkaline buffer solutions An alkaline buffer solution has a pH greater than 7. Alkaline buffer solutions are commonly made from a weak base and one of its salts. A frequently used example is a mixture of ammonia solution and ammonium chloride solution. If these were mixed in equal molar proportions, the solution would have a pH of 9.25. Again, it doesn't matter what concentrations you choose as long as they are the same How do buffers work? A buffer solution has to contain things which will remove any hydrogen ions or hydroxide ions that you might add to it - otherwise the pH will change. Acidic and alkaline buffer solutions achieve this in different ways. Acidic buffer solutions We'll take a mixture of ethanoic acid and sodium ethanoate as typical. Ethanoic acid is a weak acid, and the position of this equilibrium will be well to the left: Adding sodium ethanoate to this adds lots of extra ethanoate ions. According to Le Chatelier's Principle, that will tip the position of the equilibrium even further to the left. The solution will therefore contain these important things: lots of un-ionised ethanoic acid; lots of ethanoate ions from the sodium ethanoate; enough hydrogen ions to make the solution acidic. Other things (like water and sodium ions) which are present aren't important to the argument. Adding an acid to this buffer solution The buffer solution must remove most of the new hydrogen ions otherwise the pH would drop markedly. Hydrogen ions combine with the ethanoate ions to make ethanoic acid.

Although the reaction is reversible, since the ethanoic acid is a weak acid, most of the new hydrogen ions are removed in this way. Since most of the new hydrogen ions are removed, the pH won't change very much - but because of the equilibria involved, it will fall a little bit. Adding an alkali to this buffer solution Alkaline solutions contain hydroxide ions and the buffer solution removes most of these. This time the situation is a bit more complicated because there are two processes which can remove hydroxide ions. Removal by reacting with ethanoic acid The most likely acidic substance which a hydroxide ion is going to collide with is an ethanoic acid molecule. They will react to form ethanoate ions and water. Because most of the new hydroxide ions are removed, the pH doesn't increase very much. Removal of the hydroxide ions by reacting with hydrogen ions Remember that there are some hydrogen ions present from the ionisation of the ethanoic acid. Hydroxide ions can combine with these to make water. As soon as this happens, the equilibrium tips to replace them. This keeps on happening until most of the hydroxide ions are removed.

Again, because you have equilibria involved, not all of the hydroxide ions are removed - just most of them. The water formed re-ionises to a very small extent to give a few hydrogen ions and hydroxide ions . Alkaline buffer solutions We'll take a mixture of ammonia and ammonium chloride solutions as typical. Ammonia is a weak base, and the position of this equilibrium will be well to the left: Adding ammonium chloride to this adds lots of extra ammonium ions. According to Le Chatelier's Principle, that will tip the position of the equilibrium even further to the left. The solution will therefore contain these important things:

lots of unreacted ammonia; lots of ammonium ions from the ammonium chloride; enough hydroxide ions to make the solution alkaline. Other things (like water and chloride ions) which are present aren't important to the argument. Adding an acid to this buffer solution There are two processes which can remove the hydrogen ions that you are adding. Removal by reacting with ammonia The most likely basic substance which a hydrogen ion is going to collide with is an ammonia molecule. They will react to form ammonium ions. Most, but not all, of the hydrogen ions will be removed. The ammonium ion is weakly acidic, and so some of the hydrogen ions will be released again. Removal of the hydrogen ions by reacting with hydroxide ions Remember that there are some hydroxide ions present from the reaction between the ammonia and the water. Hydrogen ions can combine with these hydroxide ions to make water. As soon as this happens, the equilibrium tips to replace the hydroxide ions. This keeps on happening until most of the hydrogen ions are removed.

Again, because you have equilibria involved, not all of the hydrogen ions are removed - just most of them. Adding an alkali to this buffer solution The hydroxide ions from the alkali are removed by a simple reaction with ammonium ions. Because the ammonia formed is a weak base, it can react with the water and so the reaction is slightly reversible. That means that, again, most (but not all) of the the hydroxide ions are removed from the solution.

Aldehydes and Ketones


Aldehydes and ketones are simple compounds which contain a carbonyl group - a carbon-oxygen double bond. How do you name aldehydes? How do you name ketones? In aldehydes, the carbonyl group has a hydrogen atom attached to it together with a second hydrogen atom o Shown by using al o Based on alkane minus the ane e.g methane methanal

In ketones, the carbonyl group has two hydrocarbon groups attached o Shown by using one e.g propane propanone

CnH2nO Primary alcohols R-OH, when oxidized form aldehydes Secondary alcohols R-CH(OH)-R, when oxidized they form ketones The reagent is potassium dichromate(VI) K2Cr2O7, acidified with diluted sulphuric acid H2SO4 ( orange green) Aldehydes and ketones readily undergo nucleophillic attack because 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 carbonoxygen double bond very highly polar. Addition of hydrogen cyanide to aldehydes and ketones The reaction Hydrogen cyanide adds across the carbon-oxygen double bond in aldehydes and ketones to produce compounds known as hydroxynitriles. These used to be known as cyanohydrins. For example, with ethanal (an aldehyde) you get 2-hydroxypropanenitrile:

With propanone (a ketone) you get 2-hydroxy-2-methylpropanenitrile:

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. 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.

However the product is usually a 50:50 mixture of enantiomers i.e a racemic mixture Tests Add a few drops of the suspected carbonyl to bradys reagent (2,4dinitrophenylhydrazine) o a yellow orange precipitate froms

with both types of carbonyl compound o boiling points of these precipitates can be compared with a known table of values to determine its type warm a fews drops of the compound with Tollens reagent (ammoniacal silver nitrate) o only the aldehyde produces a silver mirror on the sides of the test tube o aldehydes are stronger reducing agents than ketones and reduce the metal ion Simmer with Fehlings or Benedicts solution (a blue complex of Cu2+) o The aldehyde will produce a brown/red precipitate o The ketone with stay blue Doing the triiodomethane (iodoform) reaction

Carboxylicacids
The background Carboxylic acids are compounds, which contain a -COOH group. For the purposes of this page we shall just look at compounds where the -COOH group is attached either to a hydrogen atom or to an alkyl group. Salts of carboxylic acids

Carboxylic acids are acidic because of the hydrogen in the -COOH group. When the acids form salts, this is lost and replaced by a metal. Sodium ethanoate, for example, has the structure:

Depending on whether or not you wanted to stress the ionic nature of the compound, this would be simplified to CH3COO- Na+ or just CH3COONa.

Notice: The bond between the sodium and the ethanoate is ionic. Don't draw a line between the two (implying a covalent bond). That's absolutely wrong! Although the name is written with the sodium first, the formula is always written in one of the ways shown. This is something you just have to get used to. Physical properties of carboxylic acids The physical properties (for example, boiling point and solubility) of the carboxylic acids are governed by their ability to form hydrogen bonds. Boiling points Before we look at carboxylic acids, a reminder about alcohols: The boiling points of alcohols are higher than those of alkanes of similar size because the alcohols can form hydrogen bonds with each other as well as van der Waals dispersion forces and dipole-dipole interactions. The boiling points of carboxylic acids of similar size are higher still. The higher boiling points of the carboxylic acids are still caused by hydrogen bonding, but operating in a different way. In a pure carboxylic acid, hydrogen bonding can occur between two molecules of acid to produce a dimer.

This immediately doubles the size of the molecule and so increases the van der Waals dispersion forces between one of these dimers and its neighbours - resulting in a high boiling point. Solubility in water In the presence of water, the carboxylic acids don't dimerise. Instead, hydrogen bonds are formed between water molecules and individual molecules of acid. The carboxylic acids with up to four carbon atoms will mix with water in any proportion. When you mix the two together, the energy released when the new hydrogen bonds form is much the same as is needed to break the hydrogen bonds in the pure liquids. The solubility of the bigger acids decreases very rapidly with size. This is because the longer hydrocarbon "tails" of the molecules get between water molecules and break hydrogen bonds. In this case, these broken hydrogen bonds are only replaced by much weaker van der Waals dispersion forces Making carboxylic acids Using primary alcohols/aldehydes Primary alcohols and aldehydes are normally oxidised to carboxylic acids

using potassium dichromate(VI) solution in the presence of dilute sulphuric acid. During the reaction, the potassium dichromate(VI) solution turns from orange to green. Primary alcohols are oxidised to carboxylic acids in two stages - first to an aldehyde and then to the acid. We often use simplified versions of these equations using "[O]" to represent oxygen from the oxidising agent. The formation of the aldehyde is shown by the simplified equation:

"R" is a hydrogen atom or a hydrocarbon group such as an alkyl group. The aldehyde is then oxidised further to give the carboxylic acid:

If you start with an aldehyde, you are obviously just doing this second stage. Starting from the primary alcohol, you could combine these into one single equation to give: For example, if you were converting ethanol into ethanoic acid, the simplified equation would be: The alcohol is heated under reflux with an excess of a mixture of potassium dichromate(VI) solution and dilute sulphuric acid. Heating under reflux (heating in a flask with a condenser placed vertically in it) prevents any aldehyde formed escaping before it has time to be oxidised to the carboxylic acid. By hydrolysing nitriles What are nitriles? Nitriles are compounds which contain -CN attached to a hydrocarbon group. Some common examples include:

The name is based on the total number of carbons in the longest chain including the one in the -CN group. Where you have things substituted into the chain (as in the third example), the -CN carbon counts as number 1. Where do nitriles come from? Nitriles are produced in two important reactions - both of which result in an increase in the length of the carbon chain because of the extra carbon in the -CN group.

They are formed in the reaction between halogenoalkanes (haloalkanes or alkyl halides) and cyanide ions. For example: . . . or during the reaction between aldehydes or ketones and hydrogen cyanide. For example, the reaction between ethanal and hydrogen cyanide to make 2-hydroxypropanenitrile is: Converting the nitrile into a carboxylic acid There are two ways of doing this, both of which involve reacting the carbon-nitrogen triple bond with water. This is described as hydrolysis. The two methods produce slightly different products - you just have to be careful to get this right. Acid hydrolysis The nitrile is heated under reflux with a dilute acid such as dilute hydrochloric acid. A carboxylic acid is formed. For example, starting from ethanenitrile you would get ethanoic acid. The ethanoic acid could be distilled off the mixture. Alkaline hydrolysis The nitrile is heated under reflux with an alkali such as sodium hydroxide solution. This time you wouldn't, of course, get a carboxylic acid produced - any acid formed would react with the sodium hydroxide present to give a salt. You also wouldn't get ammonium ions because they would react with sodium hydroxide to produce ammonia. Starting from ethanenitrile, you would therefore get a solution containing ethanoate ions (for example, sodium ethanoate if you used sodium hydroxide solution) and ammonia. You have to remember to convert the ions into the free carboxylic acid, because that's what we are trying to make. To liberate the weak acid, ethanoic acid, you just have to supply hydrogen ions from a strong acid such as hydrochloric acid. You add enough hydrochloric acid to the mixture to make it acidic. Now you can distill off the carboxylic acid.

Esters
Background Esters are derived from carboxylic acids. A carboxylic acid contains the -COOH group, and in an ester the hydrogen in this group is replaced by a hydrocarbon group of some kind. Usually by that of an alcohol. They are named firstly after the alcohol then the carboxylic acid, e.g:

Boiling points The small esters have boiling points, which are similar to those of aldehydes and ketones with the same number of carbon atoms. Like aldehydes and ketones, they are polar molecules and so have dipoledipole interactions as well as van der Waals dispersion forces. However, they don't form hydrogen bonds, and so their boiling points aren't anything like as high as an acid with the same number of carbon atoms. Hydrolysis of esters What is hydrolysis? Technically, hydrolysis is a reaction with water. That is exactly what happens when esters are hydrolysed by water or by dilute acids such as dilute hydrochloric acid. The alkaline hydrolysis of esters actually involves reaction with hydroxide ions, but the overall result is so similar that it is lumped together with the other two. Hydrolysis using water or dilute acid The reaction with pure water is so slow that it is never used. The reaction is catalysed by dilute acid, and so the ester is heated under reflux with a dilute acid like dilute hydrochloric acid or dilute sulphuric acid. Here are two simple examples of hydrolysis using an acid catalyst. First, hydrolysing ethyl ethanoate:

. . . and then hydrolysing methyl propanoate:

Notice that the reactions are reversible. To make the hydrolysis as complete as possible, you would have to use an excess of water. The water comes from the dilute acid, and so you would mix the ester with an

excess of dilute acid. Hydrolysis using dilute alkali This is the usual way of hydrolysing esters. The ester is heated under reflux with a dilute alkali like sodium hydroxide solution. There are two big advantages of doing this rather than using a dilute acid. The reactions are one-way rather than reversible, and the products are easier to separate. Taking the same esters as above, but using sodium hydroxide solution rather than a dilute acid: First, hydrolysing ethyl ethanoate using sodium hydroxide solution:

. . . and then hydrolysing methyl propanoate in the same way:

Notice that you get the sodium salt formed rather than the carboxylic acid itself. This mixture is relatively easy to separate. Provided you use an excess of sodium hydroxide solution, there won't be any ester left - so you don't have to worry about that. The alcohol formed can be distilled off. That's easy! If you want the acid rather than its salt, all you have to do is to add an excess of a strong acid like dilute hydrochloric acid or dilute sulphuric acid to the solution left after the first distillation. If you do this, the mixture is flooded with hydrogen ions. These are picked up by the ethanoate ions (or propanoate ions or whatever) present in the salts to make ethanoic acid (or propanoic acid, etc). Because these are weak acids, once they combine with the hydrogen ions, they tend to stay combined. The carboxylic acid can now be distilled off. Hydrolyzing complicated esters to make soap if the large esters present in animal or vegetable fats and oils are heated with concentrated sodium hydroxide solution exactly the same reaction happens as with the simple esters. A salt of a carboxylic acid is formed - in this case, the sodium salt of a big acid such as octadecanoic acid (stearic acid). These salts are the important ingredients of soap - the ones that do the cleaning. An alcohol is also produced - in this case, the more complicated alcohol, propane-1,2,3-triol (glycerol).

Because of its relationship with soap making, the alkaline hydrolysis of esters is sometimes known as saponification. Polyesters What is a polyester? A polyester is a polymer (a chain of repeating units) where the individual units are held together by ester linkages.

The diagram shows a very small bit of the polymer chain and looks pretty complicated. But it isn't very difficult to work out - and that's the best thing to do: work it out, not try to remember it. You will see how to do that in a moment. The usual name of this common polyester is poly(ethylene terephthalate). The everyday name depends on whether it is being used as a fibre or as a material for making things like bottles for soft drinks. When it is being used as a fibre to make clothes, it is often just called polyester. It may sometimes be known by a brand name like Terylene. When it is being used to make bottles, for example, it is usually called PET. Making polyesters as an example of condensation polymerisation In condensation polymerisation, when the monomers join together a small molecule gets lost. That's different from addition polymerisation which produces polymers like poly(ethene) - in that case, nothing is lost when the monomers join together. A polyester is made by a reaction involving an acid with two -COOH groups, and an alcohol with two -OH groups. In the common polyester drawn above: The acid is benzene-1,4-dicarboxylic acid (old name: terephthalic acid). The alcohol is ethane-1,2-diol (old name: ethylene glycol).

Now imagine lining these up alternately and making esters with each acid group and each alcohol group, losing a molecule of water every time an ester linkage is made.

That would produce the chain shown above (although this time written without separating out the carbon-oxygen double bond - write it whichever way you like).

Acyl (acid) chlorides

A carboxylic acid such as ethanoic acid has the structure:

There are a number of related compounds in which the -OH group in the acid is replaced by something else. Compounds like this are described as acid derivatives. Acyl chlorides (also known as acid chlorides) are one example of an acid derivative. In this case, the -OH group has been replaced by a chlorine atom.

WITH WATER: Acyl chloride + water carboxylic acid + hydrochloric acid WITH ALCOHOLS (REFLUX): Acyl chloride + alcohol ester + hydrochloric acid WITH AMMONIA: Acyl chloride + ammonia amide + hydrochloric acid WITH AMINES: Acyl chloride + amine N-substituted amide + hydrochloric acid HCl: Released as gas so steamy fumes are observed

Optical isomers
What are isomers? Isomers are molecules that have the same molecular formula, but have a different arrangement of the atoms in space. That excludes any different arrangements which are simply due to the molecule rotating as a whole, or rotating about particular bonds. Where the atoms making up the various isomers are joined up in a different order, this is known as structural isomerism. Structural isomerism is not a form of stereoisomerism, and is dealt with on a separate page Why optical isomers? Optical isomers are named like this because of their effect on plane polarised light. simple substances which show optical isomerism exist as two isomers known as enantiomers. A solution of one enantiomer rotates the plane of polarisation in a clockwise direction. This enantiomer is known as the (+) form. A solution of the other enantiomer rotates the plane of polarisation in an anti-clockwise direction. This enantiomer is known as the (-) form. So the other enantiomer of alanine is known as or (-)alanine. If the solutions are equally concentrated the amount of rotation caused by the two isomers is exactly the same - but in opposite directions. When optically active substances are made in the lab, they often occur as a 50/50 mixture of the two enantiomers. This is known as a racemic mixture or racemate. It has no effect on plane polarised light. the relationship between the enantiomers One of the enantiomers is simply a non-superimposable mirror image of the other one. In other words, if one isomer looked in a mirror, what it would see is the other one. The two isomers (the original one and its mirror image) have a different spatial arrangement, and so can't be superimposed on each other.

If an achiral molecule (one with a plane of symmetry) looked in a mirror, you would always find that by rotating the image in space, you could make the two look identical. It would be possible to superimpose the original molecule and its mirror image. Some real examples of optical isomers Butan-2-ol The asymmetric carbon atom in a compound (the one with four different groups attached) is often shown by a star.

It's extremely important to draw the isomers correctly. Draw one of them using standard bond notation to show the 3-dimensional arrangement around the asymmetric carbon atom. Then draw the mirror to show the examiner that you know what you are doing, and then the mirror image.

You might also like