WBBSE Notes For Class 8 Geography Chapter 5 Cloud And Rain

Chapter 5 Cloud And Rain

What is Cloud: Clouds, which are condensed water vapour, are one of the most visible marks (elements) of the weather. They form in distinctive patterns and often give a quick clue as to what weather might happen in the near future. Meteorologically clouds are very significant because all forms of precipitation occur from them. It may be mentioned that not all clouds yield precipitation but no precipitation is possible without clouds.

Read and Learn all WBBSE Notes For 8 Class Middle School Geography

Chapter 5 Cloud And Rain: The Cloud Family

 

Clouds can be classified in two ways—
(1)According to height and
(2)According to shape and formation.

(1) Classification according to height: By height, clouds can be divided into three types:
1. High clouds
2. Medium clouds
3. Low clouds

(2) Classification according to shape and formation: Clouds can be classified into four main types which are given Latin names. They are
1. Cirrus—Looks like a feather or fine scales,
2. Stratus—they have a layered structure
3. Cumulus—they look like a mound of fully cotton and
4. Nimbus—they are rain clouds, giving plenty of rainfall.

 

WBBSE Notes For Class 8 Geography Chapter 5 Cloud And Rain The Cloud Family

 

High clouds: The lower limit of this cloud is 6000 to 12000 metres or above 20000 feet. There are mainly three types of clouds in the High cloud family. These are as follows:

1. Circus: 3 The high altitude detached clouds having feather-like or fibrous appearance are called ‘cirrus clouds’ They are composed of tiny ice crystals and are transparent and white in colour but have brilliant colour at sunset and sunrise. When the weather is clear these clouds are seen and-no rain occurs. So this cloud generally indicates fair weather.

 

WBBSE Notes For Class 8 Geography Chapter 5 Cloud And Rain Cirrus Cloud

2. Cirrocumulus: Cirrocumulus clouds are white-coloured clouds having cirri form layers or wave-like forms. They generally appear as the scaled body of a mackerel fish. So, the cirrocumulus cloud ‘Mackerel Sky’.This cloud usually indicates fair weather because from these clouds rain occurs.

3. Cirrostratus: Cirrostratus clouds are generally white in colour and spread in the sky like milky thin sheets. The sun and moon seen through this cloud have a halo. This cloud forms a rainbow. It signals storms.

WBBSE Notes For Class 8 Geography Chapter 5 Cloud And Rain Cirro-stratus Cloud

Medium clouds: The lower limit of this cloud in between 2100 mts to 6000 metres (6500 to 20000 feet). There are mainly two types of clouds in the Medium Cloud falily. These are:
(1)Alto-cumulus and
(2)Alto-stratus

(1) Alto-cumulus: Alto-cumulus clouds are characterized by wavy layers of globular form. Sometimes these clouds are called as ‘sheep clouds’ or ‘wool- pack clouds because they are looking like scattered mounds of wool. They appear white or grey in colour. In between the clouds blue sky is seen These clouds generally herald clear weather but when they accumulate in large amounts rain may occur.

WBBSE Notes For Class 8 Geography Chapter 5 Cloud And Rain Alto-cumulus Cloudv

(2) Alto-stratus: Alto-stratus clouds are thin sheets of grey or blue colour having fibrous or uniform appearance. The sun looks very hazy through this cloud. Rain occurs for a long duration and over large areas from this cloud.

WBBSE Notes For Class 8 Geography Chapter 5 Cloud And Rain Altro-stratus Cloud

Low clouds: The lower limit of this cloud is less than 2100 m. or 6500 feet. Most clouds are stratus or layered. There are mainly three types of clouds in the Low clouds family. These are
(1) Strato cumulus,
(2) Stratusand
(3) Nimbostratus.

(1) Strato-cumulus: Strato-cumulus clouds are found in rounded patches between the height of 2500 m to 3000 m. They are of grey or whitish clour. They are composed of globular masses or rolls which are generally arranged in lines, waves or groups. So, they are also called ‘Bumpy clouds’.

They are generally associated with fair or clear weather but occasional rain on snow is not ruled out. During winter in temperate regions, often the sky is covered by this cloud giving heavy rainfall.

WBBSE Notes For Class 8 Geography Chapter 5 Cloud And Rain Strato-cumulus Cloud

(2) Stratus: Stratus clouds are dense, low-lying fog-like clouds of dark grey colour but are seldom close to the ground surface. They are composed of several uniform layers and cover the entire sky. It does not extend high from the surface so creates invisibility at low levels. It is a great problem for mountaineers and air pilots. Often this cloud occurs drizzle.’

(3) Nimbostratus: Nimbostratus clouds are low clouds of dark grey to black colour, very close to the ground surface. They are so compact and thick (hundreds of metres) that there is complete darkness and there is copious precipitation. ‘Nimbo’ is derived from the Latin word ‘nimbo’ meaning thereby ‘rainstorm’, it indicates bad weather because it causes heavy rainfall. During cyclones it covers the whole sky and gives rainfall.

WBBSE Notes For Class 8 Geography Chapter 5 Cloud And Rain Nimbo-stratus Cloud

Vertical clouds or Clouds of great vertical extent: The height of the clouds extend from 1500 m to 9000 m, resembling a tree from below to above. There are mainly two types of vertical clouds. These are—
(1)Cumulus and
(2)Cumulonimbus.

(1) Cumulus: Cumulus clouds are very dense, widespread, dome-shaped and have flat bases. They start from low levels but reach high above. The bottom of the cloud is flat but its top resembles a cauliflower. The lower part of the cloud is dark grey or black but the higher part is white and the sun looks bright through it. It usually indicates clear and fair weather.

WBBSE Notes For Class 8 Geography Chapter 5 Cloud And Rain Cumulus Cloud

(2) Cumulonimbus: Cumulonimbus clouds are thunderstorm clouds. It is a huge dense cloud. The bottom and middle of the cloud is black in colour, the sides are white or grey.

It has a great vertical extent of about 4000 m upward from the layer of air adjacent to the earth’s surface. During ‘Kalbaisakhi’ these clouds are seen in the North-west corner of the sky. These clouds cause heavy rain, storms or hailstorms accompanied by lightning, thunder and gusty winds. So it-is also called ‘Thunder Cloud’.

WBBSE Notes For Class 8 Geography Chapter 5 Cloud And Rain Cumulo-nimbus Cloud

The origin of clouds: Cloud is the most common form of ‘Condensation’. When moist air rises, it cools down. The water vapour present in the air also cools and changed into minute droplets of water or crystals. These droplets of water or ice crystals stick to dust and salt particles, floating in the atmosphere. These small water droplets or ice crystals form a floating mass called cloud.

Thus, a cloud is nothing but a collection of countless droplets of water or the tiniest crystals of ice suspended in the atmosphere. These droplets or crystals are so small that they are blown about and carried with the slightest movement of air.

WBBSE Notes For Class 8 Geography Chapter 5 Cloud And Rain Origin Of Clouds

Evaporation: The process where by a liquid (water) changes into a gas (water vapour) is called evaporation. The water of the water bodies is changed into water vapour by solar energy. It mixes with other gases of the atmosphere. At 10°C. temperature, on cubic metre air can hold 11-4 gm water vapour. Warm air can absorb more water vapour than cold air.

Saturated air: When a specific quantity of airmass, at a particular temperature holds the maximum amount of water vapour as it can possibly hold is called ‘saturated air’. The amount of water vapour required to saturate the air depends on temperature and pressure.

Dew point: The temperature of air at which it becomes saturated with water vapour starts to condense to form water droplets is called ‘dew point’. The upper level of saturation is known as dew point.

Condensation: The physical process of transformation from the vapour to the liquid state is called ‘condensation’. In the atmosphere, condensation occurs either when the temperature drops sufficiently for moisture to be cooled to its dew point, or when there is enough water vapour within an airmass for it to reach saturation point.

Distinguish between Evaporation and Condensation:

Evaporation Condensation
1. Evaporation is a process by which water changes into a vapour. 1. Condensation is a process by which water vapour present in the air changes into minute droplets of water.
2. It takes place more when the weather is hot, dry and windy. 2.  It takes place when the temperature of the saturated air goes down below the dew point.
3. It also takes place more quickly in the lower latitudes. 3.  It also takes place more quickly at higher latitudes than at lower latitudes.
4. As a result of evaporation the amount of water on the surface of the earth goes on decreasing and decreasing 4. As a result of our censer or the amount of water goes on increasing arc increasing.

 

Some forms of condensation:
(1) Fog: Droplets of water suspended in the lower layers of the atmosphere resulting from the condensation of water vapour around nuclei of floating dust or smoke particles is called ‘fog’. Its visibility is less than 1 km.

WBBSE Notes For Class 8 Geography Chapter 5 Cloud And Rain Fog

(2) Mist: A reduction of visibility within the lower atmosphere to 1—2 km caused by condensation producing v/ater droplets within the lower layers of the atmosphere is called ‘mist’. It is intermediate between fog and haze.

WBBSE Notes For Class 8 Geography Chapter 5 Cloud And Rain Mist

 

(3) Haze: An obscurity of the lower atmosphere that limits visibility to under 2 km, but over 1 km, is called ‘haze’. It is normally formed by water particles that have condensed around nuclei to the atmosphere but may also be a result of particles of smoke, dust or salt in the air.

 

WBBSE Notes For Class 8 Geography Chapter 5 Cloud And Rain Haze

 

(4) Smog: A form of fog that occurs in areas where the air contains a large amount of smoke is called ‘smog’ (Smoke + fog).

WBBSE Notes For Class 8 Geography Chapter 5 Cloud And Rain Smog

Humidity: Humidity means the amount of water vapour present in the air. Water vapour is always present in the air. About 2% of the atmosphere consists of water vapour. Most of the water vapour of the atmosphere comes from the oceans, lakes, and rivers through evaporation.

(1) Absolute Humidity: Total amount of water vapour present in air at a particular temperature is absolute humidity. It is defined as the weight of water vapour per unit volume of air. It is expressed as grams per cubic metre of air.

Distribution:
1)Absolute humidity is maximum in the equatorial region,
2)It is more in summer than in winter, similarly, it is higher during the day than at night.

(2) Relative Humidity: Relative humidity is expressed as a percentage. It is the ratio between the actual amount of the water vapour present in the air at a temperature and the maximum amount of water vapour that the same volume of air can hold at a given temperature.

 

WBBSE Notes For Class 8 Geography Chapter 5 Cloud And Rain

In other words, it is a ratio, expressed in percentage, between the absolute humidity and the maximum vapour capacity of the air.

WBBSE Notes For Class 8 Geography Chapter 5 Cloud And Rain

Example: An air at 20°C temperature has 8 grams of water vapour per cubic metre actually present. But the air at 20°C temperature can hold 16 (gms) of water vapour as per cubic metre.

E:\websites-Haritha\west bengal.guide\Geography Class 8\images\chapter 5\WBBSE Notes For Class 8 Geography Chapter 5 Cloud And Rain.png

When the relative humidity is 100%, the air is fully saturated.

Distribution:
1) Relative humidity is maximum at the Equator and decreases polewards,
2) It is low in hot deserts, in continental interiors, as in areas of anti-cyclonic conditions.

3)Specific Humidity: The humidity of the atmosphere expressed as the ratio of the weight of water vapour (in grams) to the total weight (in kilograms) of a given volume of air is called Specific humidity. This varies from about 0-2 gm/Kg in very dry cold arctic air to over 18-0 gm/Kg in hot humid tropical air.

Distinguish between Relative humidity and Absolute Humidity:

Relative Humidity Absolute Humidity
1. It is a ratio between the actual amount of water vapour present In the air at a particular temperature and the maximum amount of water vapour that the air can hold at that temperature. 1. The total amount of water vapour present in the air at a particular temperature is absolute humidity.
2. It is a ratio between the absolute humidity and vapour capacity of the air. 2. It is defined as the weight of water vapour per unit volume of water.
3. It is maximum In the equatorial region, but is lowest in hot deserts. 3. It Is maximum over oceans and lowest in high-pressure areas.
4. It is expressed as a percentage. 4.  It is expressed as grams per cubic metre of air.

 

The Importance of Relative Humidity: Relative humidity is daily life. Some important points have been mentioned here.
(1) It is possible to find out the possibility of rainfall with the help of relative humidity.
(2) Relative humidity is essential to agricultural operations.
(3) Many chemicals and medicines deteriorate in high relative humidity.
(4) It is necessary to take into account the relative humidity before constructing buildings.
(5) Radio, Television, electric instruments and other scientific instruments are affected by relative humidity.
(6) High relative humidity do not enjoy good health. Even very low relative humidity is not conducive to health. It has been found that a relative humidity of 60% is most suitable for human health.

WBBSE Notes For Class 8 Geography Chapter 5 Cloud And Rain Hygrometer

 

The Measurement of Relative Humidity: Relative humidity is measured with the help of various types of Hygrometers. The most common Hygrometer is known as ‘Dry and West Bulb Thermometer’.

Precipitation: The particles of water or ice that form within clouds and fall towards the earth surface is Precipitation. According to famous climatologist H. J. Critehfield, “Precipitation is defined as water in liquid or solid forms falling to the earth.” Precipitation includes all forms in which moisture falls on the earth’s surface.

It is the process by which condensed water from clouds falls on the earth’s surface. It can be solid or liquid. Precipitation is a complex process when millions of drops of water combine together and fall on the earth.

Forms of Precipitation: Different forms of precipitation are:
1. Rainfall,
2. Snowfall,
3. Hail,
4. Dew,
5. Frost,
6. Sleet,
7. Drizzle,
8. Glaze.

1. Rainfall: The release of moisture in the form of drops of water is called ‘rainfall’.After condensation of moist laden air, clouds are formed. Teh drops of water become so heavy that air cannot hold these. The falling of these drops of water from the clouds is called ‘rainfall’.

2. Snowfall: A from of precipitation consisting of crystals of ice is called ‘snow’.it is produced when condensation takes place at a temperature below freezing point, (0°C), so that the minute crystals of ice form directly from the water vapour. These small crystals then unite to form flakes of ice called snow.The coming down of snow flakes towards the earth’s surface is known as ‘snowfall’.It is common in higher latitudes and high mountain regions.

WBBSE Notes For Class 8 Geography Chapter 5 Cloud And Rain Snowfall

 

3. Hall: Precipitation in the form of pellets of ice (hailstone) that develop in and fall from a cumulonimbus cloud, either at a cold front or where intense heating of the surface causes rapidly ascending convection currents is called hail. When they become so big and heavy that the air can no longer hold them. Therefore, they fall back to the earth’s surface as hailstones. Hailstones cause damage to crops and buildings.

WBBSE Notes For Class 8 Geography Chapter 5 Cloud And Rain Hailstrom

 

4. Dew: When the temperature for the air is higher than the dew point and a cool object having a lower temperature comes in contact with such air. This causes condensation in the air The drops of water thus formed rest on the cold object like grass, leaves, rocks, etc. These drops of water on the cold objects are called dew. The conditions favouring dew formations are moist air, light winds, and clear night skies to ensure maximum cooling by radiation.

WBBSE Notes For Class 8 Geography Chapter 5 Cloud And Rain Dew

 

5. Frost: A weather condition that occurs when the air temperature is at or below the freezing point (0°C), the water vapour will condense to form an icy deposit. It is called frost.

6. Sleet: In the U. K., Sleet is described as a form of precipitation consisting of either partly melted snow-flakes or rain and snow falling together. In the.U.S.A., Sleet is described as a form of precipitation consisting of frozen raindrops that have subsequently partially remelted. So Sleet is considered as a mixture of rain and snow.

7. Drizzle: Light rainfall, consisting of particles of size less than 0-5 mm in diameter and are close together is called ‘drizzle’. Normally drizzle is produced by stratus and strato- cumulus clouds.

8. Glaze: A covering of smooth clear ice that coats objects and surfaces is called ‘glaze’. It occurs when supercooled water droplets freeze on contact with a surface and when a fall in temperature causes wet surface to freeze.

Distinguish between Precipitation and Rainfall:

Precipitation Rainfall
(1) The particles of water or ice that form within clouds and fall towards the earth is precipitation. (1)The release of moisture in the form of drops of water is called rainfall.
(2) It is the collective name given to different forms of moisture after condensation. (2)It is a type of precipitation when moisture falls on the earth in the form of droplets of water.
(3) Precipitation has two forms—liquid and solid. (3)Rainfall has only liquid form.
(4) Rainfall, snow, hail, drizzle, sleet, dew, frost and glaze are the common forms of precipitation. (4) Three major types of rainfall are—(1)Convectional,(2)Relief or Orographic and (3)Frontal or cyclonic rainfall.

 

How is Rainfall caused: Humidity is the basis of rain. The general cause of rainfall is the cooling of saturated air. The mechanism of rainfall passes through a number of stages:
(1)Air should have plenty of water vapour,
(2)That air should be saturated with moisture,
(3)The air should get cooled or get the chance to come in contact with the cold air for condensation and
(4)Condensation of water vapour makes droplets. Thus the release of moisture in the form of drops of water is called ‘rainfall’.

Types of Rainfall: On the basis of mode of formation and process of cooling of air, there are three types of rain.
(1) When the air is cooled by rising convection currents (Convectional rainfall).
(2) When moist air is cooled due to ascent along a mountain barrier (Orographic or Relief rain).
(3) When air is cooled due to the meeting of warm and cold air masses.(Frontal or Cyclonic rain).

(1) Convectional Rainfall: Convectional rainfall is formed when the air is cooled by the rise of the convectional current. It happens when the land is intensely heated, the hot and moist air rises up vertically as convection currents. As it rises up, it expands and cools due to the release of pressure. Hence, condensation takes place resulting in a cumulus or cumulonimbus clouds.

Consequently, it results in torrential rainfall. The equatorial region of low latitude experiences convectional rainfall. In this region, the convectional system is best developed because daily heating of the ground surface up to noon causes convection currents. Consequently, the sky becomes overcast by 2—3 p.m. daily causing pitch darkness and heavy rains and the sky becomes clear by 4 p.m. So it is called 4 O’ clock rain. Thus, the convectional rainfall in the equatorial region is a daily regular feature.

WBBSE Notes For Class 8 Geography Chapter 5 Cloud And Rain Convectional Rainfall

 

Characteristic features of convectional rainfall:
(1) It occurs daily in the afternoon in the equatorial region.
(2) It is of very short duration but occurs in the form to heavy showers.
(3) It occurs through thick dark and extensive cumulonimbus clouds.
(4) It is accompanied by cloud thunder and lighting.

(2) Orographic or Relief Rainfall: Orographic rainfall occurs when moist air is cooled due to ascending along a mountain slope or a plateau barrier. The moist wind coming from the sea rises up along the slope of the land and comes in contact with cool upper atmosphere.

As a result the moist air condenses and rainfall occurs. Since this type of rainfall is caused by the relief of the land, it Is known as relief rainfall’. The the windward slopes’ gets maximum rainfall.

WBBSE Notes For Class 8 Geography Chapter 5 Cloud And Rain Relief Rainfall

 

Distinguish between “Leward slope’ and ‘Windward slope”:

Windward Slope Leward Slope
(1) Windward side is the side of the mountain which faces the moisture (rain-bearing winds). (1) The Leward slope/side is the side of the mountain which is opposite the windward side.
(2) On the windward side, the air ascends and becomes cool. (2) On the Leeward side the air descends and gets warm.
(3) When the moisture-bearing winds climb the windward side of the mountain, they cool down and bring heavy rainfall. (3) When these air cross over to the other side (Leward side) they have already lost much of their moisture. While descending they further become warm and dry so they give less rainfall on the Leeward side.
(4) The western slope of the Western Ghats gets more rain because they form the windward side. (4) The Eastern slopes of the Western Ghats get lesser rain because they form the Leeward side.

Chapter 5 Cloud And Rain: Characteristic Features Of Orographic Rainfall

 

  • The windward slope, also called as rain slope, receives the maximum amount of rainfall whereas the leeward side of the mountain gets vey low rainfall. .
  • There is maximum rainfall near the mountain slopes and it decreases away from the foothills.
  • Orographic rainfall may occurs in any season’ Unlike other types of rainfall it is more widespread and of long duration.
  • The amount of rainfall increases with increasing height along the windward slopes of the mountains.

Distinguish between Convectional Rainfall and Orographic Rainfall:

Convectional Rainfall Relief/Orographic Rainfall
1. This rainfall is caused by convection currents. 1. This rainfall is caused when a mountain forces the winds to rise.
2. The rising air expands and is cooled to give rainfall. 2. It occurs in mountainous regions.
3. It gives heavy showers for a short period. 3. It gives heavy rainfall on a windward slope, but Leeward slope is dry and is called Rainshadow
4. The equatorial region gets heavy showers in the afternoon daily. 4. South-West Monsoons give heavy rainfall on the Western Ghats, but the Deccan plateau lies in rain-shadow.

 

(3) Cyclonic or Frontal Rain: When hot and air masses move towards each other, parallel to the earth’s surface along a front, the cold air being heavy flows near the earth’s surface, while the hot air blows above it, rises and expands to cool and condense to form rainfall.

This is called “Cyclonic Rain’ because as this rain is associated with cyclones. This rain is common in North-West Europe in winter, Punjab plains experience rainfall in winter from cyclones coming from the Mediterranean Sea.

WBBSE Notes For Class 8 Geography Chapter 5 Cloud And Rain Cyclonic Or Frontal Rainfall

Chapter 5 Cloud And Rain: Characteristics of Cyclonic Frontal Rain

  1. This rainfall is slow, continuous and extensive.
    The mechanism of cyclonic rainfall is of two types:
    1)Temperate cyclone and 2) Tropical cyclone.
  2. This rainfall associated with a warm front is widespread and long-duration.
  3. Sometimes this rainfall- occurs in the form of snowfall and hailstorms.
  4. Most of the rains of temperate regions are received through cyclones.
  5. If tropical cyclones are full of moisture become saturated and yield heavy showers characterized by lightning and thunder.

Raingauge: The amount of rainfall is measured by an instrument called ‘Raingauge’.

WBBSE Notes For Class 8 Geography Chapter 5 Cloud And Rain Raingauge

Importance of Precipitation: Precipitation is an important factor of weather and climate. Rainfall often brings a cooling effect on the weather. During intense summer rainfall offers a great relief of the people. Temperature goes down due to heavy showers. Snowfall often brings very chilled weather and people suffer from shivering cold.

WBBSE Notes For 8 Class Middle School Geography

Equilibrium – Definition and Types

Equilibrium

You have learnt about several aspects of a chemical reaction so far. You know, for example, that certain proportions of reactants react to form particular proportions of products and that these proportions are specified in a chemical equation. In the previous chapter, you learnt about the energy changes accompanying a chemical reaction. But there is one aspect of a chemical reaction that we have not considered at all.

Suppose we take reactants in exactly the required proportions (in the ratio of their respective number of moles in the balanced equation) and mix them together. Is it necessary that the reaction will proceed to completion? In other words, is it necessary that all of the reactants will be converted into products?

  • We assume this in stoichiometric calculations (chemical arithmetic), but this is not always true. Quite often, a chemical reaction stops after a while and the resultant mixture contains reactants and products. It is said that the reaction has reached a state of chemical equilibrium.
  • This state remains unchanged and the composition of the mixture remains the same, unless something forces a change. This something could be a change in temperature, pressure, or concentration of the reactants.
  • It could even be a change in the concentration of the products, which could be achieved by removing the products (already formed) from the reaction site. These considerations, i.e., how to change the state of chemical equilibrium, are very important for industrial chemists. You can well imagine that the productivity of an industrial process may depend on these factors.
  • When a reaction system reaches a chemical equilibrium, all the properties of the system, i.e., temperature, pressure, concentration of reactants and concentration of products, remain constant over time. What actually happens in such a system is that two opposing processes take place at the same time. Just as the reactants react to form the products, the products combine to produce the reactants.

And there comes a time when the forward process and the reverse process occur in such a way that the reaction seems to come to a standstill. At equilibrium, the rates at which the two opposing processes occur become equal.

  • A double arrow (\(\rightleftharpoons\)) is used to denote an equilibrium, whereas a single arrow (-») is used to denote a reaction which proceeds to completion. The extent to which the chemical reactions proceed may be different. Some reactions proceed nearly to completion and only a negligible concentration of the reactants is left.
  • There are reactions in which the concentrations of reactants and products are comparable at equilibrium. Lastly, there are a few reactions in which most of the reactants remain unchanged at equilibrium, i.e., small amounts of products are formed.
  • Our discussion so far has been centred around chemical reactions, so you may be tempted to believe that a state of equilibrium can be reached only in chemical processes. But this is not true.
  • A state of equilibrium may be reached in a physical process as well. When the forward process and the reverse process occur simultaneously at exactly the same rate, a state of equilibrium is reached.

Equilibrium In Physical Processes

You have already come across physical processes which reach a state of equilibrium. A liquid in equilibrium with its own vapour inside a closed vessel is a system in which the processes of evaporation and condensation proceed at the same rate. The examples of physical equilibrium are solid \(\rightleftharpoons\)liquid, liquid \(\rightleftharpoons\) gas, solid \(\rightleftharpoons\) solution and gas solution.

Solid-liquid equilibrium: If you drop a few ice cubes into a glass of water, all the ice changes into water after some time. This physical process or change proceeds in a particular direction until it is complete, i.e., until all the ice melts into water. Now suppose you pour a mixture of ice cubes and water at 273 K and normal atmospheric pressure into a perfectly insulated flask.

  • The system, comprising the mixture of ice and water in the flask, will not exchange heat with the surroundings and a state of equilibrium will be reached, in which neither the temperature, the pressure, nor the composition of the mixture will change with time.
  • It is not as though nothing will happen inside the flask, but there will be no way of telling that something is happening since there will be no outward indication.
  • If you could become really small and enter the world of the molecules of water and ice, you would be able to see that what really happens is that molecules of ice go into the liquid state and molecules of water collide and stick to ice, constantly.

The two processes go on simultaneously and at the same rate. Consequently, neither the mass of ice nor the mass of water changes.

To put this in terms of change in free energy, when an ice-water system is in equilibrium at 273 K and at 1 atm pressure, ΔG = 0. For the process, ice \(\rightleftharpoons\) water

ΔG < 0 when T > 273 K, and

ΔG > 0 when T < 273 K.

  • The system comprising ice and water at 1 atm pressure can be at equilibrium only at 273 K. The solid and liquid phases of any (pure) substance at 1 atm pressure can be in a state of equilibrium only at a particular temperature.
  • This temperature at which the solid and liquid phases of a pure substance are at equilibrium at a pressure of 1 atm is called the normal melting point or the normal freezing point of the substance.

The same definition but at a pressure of 1 bar refers to standard freezing point of a substance. The equilibrium between the solid and liquid phases of a substance at its melting point is a dynamic equilibrium, in which the activity in one direction is balanced by the activity in the reverse direction. Let us recall the characteristics of a system in dynamic equilibrium.

  1. The forward and reverse changes (processes) occur simultaneously and at the same rate, so that there is no change of mass of either side of the equilibrium.
  2.  ΔG = 0

Liquid-vapour equilibrium: You have already come across this kind of equilibrium while studying vapour pressure. It can be demonstrated easily by performing a simple experiment. Take a small quantity of water at room temperature in a closed vessel connected to a manometer.

  • Take care to see that the vessel has been evacuated first. To begin with, the level of mercury in both limbs of the manometer will be the ⇒ same. Then the level of mercury in the right limb of the manometer will rise slowly and the level of mercury in the left limb will fall.
  • The level of water in the vessel falls gradually. There comes a time when the levels of mercury in tire two limbs become constant and so does the level of water in the vessel.

Basic Chemistry Class 11 Chapter 7 Equilibrium Experiment To Demonstrate Liquid Vapour Equilibrium

  • Initially, there is no water vapour in the vessel and the level of mercury in both limbs is the same. As the water starts evaporating and more and more molecules escape into the gaseous phase, the pressure exerted by the water vapour increases, leading to a drop in the mercury level in the left limb and a rise in the right limb.
  • But with the accumulation of vapour molecules in the vessel, the reverse process (condensation) starts. Some vapour molecules revert to the liquid phase. Condensation occurs at a lower rate than evaporation, to start with. But after a while, the two rates become equal and a state of equilibrium is established.

The pressure exerted by the vapour molecules becomes constant, as shown by the steady level of mercury in the manometer. The level of water in the vessel becomes constant because there is no net evaporation. The number of molecules entering the gas phase at any time is the same as the number of molecules entering the liquid phase.

∴ \(\mathrm{H}_2 \mathrm{O}(\mathrm{l}) \rightleftharpoons \mathrm{H}_2 \mathrm{O}(\mathrm{g})\)

The pressure exerted by the water vapour at equilibrium is called the equilibrium vapour pressure or simply the vapour pressure of water at the given temperature. In general, the pressure exerted by the vapour of a liquid in equilibrium with the liquid at a particular temperature is called the vapour pressure of the liquid at that temperature.

  • Vapour pressure changes with temperature but is independent of the volume of the liquid. It does not depend on the volume of the container either.
  • Different liquids have different vapour pressures at the same temperature and the liquid which has a higher vapour pressure is more volatile, or boils at a lower temperature. For instance, if the vapour pressure of liquid A is 3.2 kPa and that of another liquid B is 10.3 kPa at the same temperature, B is more volatile than A, or B has a lower boiling point than A.
  • This should not be difficult to understand if you bear in mind that the vapour pressure of a liquid rises with temperature and that its boiling point is that temperature at which its vapour pressure becomes equal to the atmospheric pressure.
  • Consider the two liquids A and B. The vapour pressure of B is higher than that of A at a particular temperature. As the temperature rises, the vapour pressures of both the liquids will rise.

Obviously, the vapour pressure of B will become equal to the atmospheric pressure at a temperature lower than that at which the vapour pressure of A will become equal to the atmospheric pressure. Consequently, the boiling point of B will be lower than that of A.

Solid-vapour equilibrium: In case of sublimable solids, the particles may escape occasionally into the vapour phase and establish vapour pressure.

  • The particles which are at the surface of the solid and which have, at any point of time, higher average kinetic energy than the others, may escape into the vapour phase. If the solid is kept in a closed cylinder, particles are not able to escape and add to the vapour phase.
  • Eventually, a point is reached when the particles from the vapour phase start returning to the solid phase. As in the case of liquid-vapour equilibrium initially, this reverse process is slow but after some time, an equilibrium is reached whereby rate of return of the particles to the solid phase becomes equal to the rate of escape.
  • A dynamic equilibrium is thus established. The vapour pressure now corresponds to the equilibrium vapour pressure, which is characteristic of the solid.
  • The tendency of the particles of the solid to escape depends on the intermolecular forces. Tire larger the intermolecular interaction, the less will be the vapour pressure of that solid.

Temperature also has a considerable effect on the vapour pressure of solids. The higher the temperature, the more energetic are the particles and the higher is the vapour pressure.

Basic Chemistry Class 11 Chapter 7 Equilibrium Experiment To Demonstrate The Dynamic Equilibrium Between The Solute In Solid State And The Solute In Solution

Iodine is a volatile solid. When solid iodine is placed in a closed vessel and the vessel is heated, the region above the solid becomes violet due to iodine vapours. The equilibrium is represented as \(\mathrm{I}_2 \text { (solid) } \rightleftharpoons \mathrm{I}_2 \text { (vapour). }\)

Another example is dry ice (solid CO2); it sublimes directly from solid to gas at atmospheric pressure.

  • So far we have been considering closed systems. Systems we encounter in our everyday lives are (generally) not closed. The atmosphere, into which water vapour from lakes and rivers and other water bodies escapes, is an open system.
  • Though the rate of evaporation of a liquid at a particular temperature in an open system is the same as that in a closed system, equilibrium can never be reached in an open system.
  • This is because the rate of condensation can never be equal to the rate of evaporation in an open system, in which the vapour molecules get dispersed in a large volume.
  • The water molecules which escape into the atmosphere, for instance, do not remain confined to a small space, they, get dispersed by the wind.
  • How much water vapour is there in the air (what we refer to as humidity) depends on several factors, such as the presence of water bodies in the area, the temperature and the velocity of the wind.
  • At places close to the sea, especially when the wind velocity is not high, the accumulation of water vapour in the air leads to acute discomfort.

Solid-solution equilibrium: It is true that sugar dissolves in water. But you cannot dissolve an indefinite amount of sugar in a particular amount of water. You can, of course, dissolve more sugar in a particular amount of water if you heat the solution.

  • This is what they do in sweet shops when they prepare sugar syrups. But when the syrup cools, crystals of sugar separate from the solution because when the solution (syrup) cools, the solubility of the solute (sugar) in the solvent (water) decreases.
  • You may remember from your science courses in junior classes that a solution in which more solute cannot be dissolved is called a saturated solution.
  • The amount of a solute that has to be dissolved in a certain amount of a solvent to prepare a saturated solution at a particular temperature is called the solubility of the solute in that solvent at that temperature.
  • Solubility is usually expressed in terms of the number of grams of a solute that dissolves in 100 mL of the solvent.
  • The solubility of a solid in a liquid depends on temperature, but we will come to that later. Pressure has hardly any effect on the solubility of a solid in a liquid.

What happens when a solution becomes saturated is that a dynamic equilibrium is established between the molecules of the solute in the solid state and the molecules of the solute in the solution.

Solute (in solution) \(\rightleftharpoons solute\) (solid)

  • The process by which molecules of the solid solute go into solution is called dissolution. The reverse process by which molecules of the solute in the solution revert to the solid state is called precipitation.
  • When a solute is added to a solvent, there are no molecules of the solute in the solvent to begin with. Slowly, as the solute dissolves in the solvent, the reverse process begins.

At first, the process of dissolution proceeds at a faster rate than the process of precipitation. Than there comes a point when the two processes proceed at the same rate. This happens when equilibrium is reached.

Rate of dissolution = rate of precipitation

It is not as though activity comes to a standstill in the solution, but the number of molecules of the solute going into the solution (in a certain time) becomes the same as the number of molecules returning to the solid state.

  • This can be demonstrated with the help of an experiment. Prepare a saturated sugar solution in a beaker in such a way that some undissolved sugar remains at the bottom of the beaker.
  • If you add some sugar-containing radioactive carbon to the solution, in a while the solution and the undissolved sugar contains sugar which has radioactive carbon.
  • This shows that though the solution is saturated, the process of dissolution continues and that the rate of dissolution must be equal to the rate of precipitation, since the amount of sugar left undissolved remains the same.
  • In other words, the experiment demonstrates the dynamic nature of the equilibrium reached.

Gas-solution equilibrium: You must have heard people referring to drinks like Pepsi or Mirinda as carbonated or aerated water.

  • Well, strictly speaking, such drinks also contain sugar and other additives and it would be more correct to refer to soda as aerated water.
  • Nonetheless, it is true that drinks like Pepsi, too, have carbon dioxide dissolved under pressure. It is this carbon dioxide which escapes with a hissing sound when a bottle of coke or soda is opened.
  • The fizz that you enjoy in such drinks is also because of the dissolved carbon dioxide.
  • Aerated drinks are manufactured by dissolving carbon dioxide in them at a high pressure. Unlike solids, the solubility of gases (in liquids) increases with pressure.

The effect of pressure on the solubility of a gas in a liquid is given by Henry’s law (William Henry was a British physician and chemist), which says that the mass of a gas dissolved in a given mass of a solvent at a particular temperature is proportional to the pressure of the gas above the solvent.

  • The solubility of a gas in a liquid decreases as temperature increases. When an aerated drink is bottled there is a dynamic equilibrium between the gas molecules above the solution (tire drink) and the gas molecules in the solution.
  • This equilibrium holds as long as the pressure is maintained. When the bottle is opened, the pressure above the solution falls and some of the dissolved gas escapes to reach a new equilibrium.
  • If an aerated drink is left open to the air for some time, it turns ‘flat’ because most of the dissolved gas escapes as the pressure above the drink (solution) becomes equal to the atmospheric pressure.

Henry’s law can be expressed mathematically as follows.

∴ m∝ p or m = kp,

where m is the mass of the gas dissolved in the solution, p is the pressure of the gas above the solution and k is a constant of proportionality known as the Henry constant.

  • Let us look at this another way. The mass of the gas dissolved in the solution will be related to the concentration of the gas in the solution.
  • The pressure exerted by the gas above the solution will depend on the concentration of the gas above the solution. This gives us another way of expressing Henry’s law.
  • Concentration of gas in solution x concentration of gas above solution

or \(\frac{\text { concentration of gas in solution }}{\text { concentration of gas in gaseous phase }}=\text { constant. }\)

Example: Suppose 0.30 g of iodine is stirred in 100 mL of water at 288 K until equilibrium is reached. Hoiv much iodine would dissolve in the water, if the solubility of iodine in water at 288 K is 0.0011 mol L-1 [Another way of expressing this is I2 (aq) at equilibrium = 0.0011 mol L-1 at 288 K]. Suppose another 100 mL of water is added to the solution after equilibrium is reached with 0.30 g of iodine and 100 mL of water. How much iodine would be left undissolved and what would be the concentration of iodine in the solution?
Solution:

The solubility of iodine (concentration at equilibrium) is given as 0.0011 mol L-1 at 288 K.

This means at equilibrium mass of iodine dissolved in 1 L of water = 0.0011 x 254 ≅ 0.279 g s 0.28 g [molar mass of I2 = 254 g]

∴ mass of iodine dissolved in 100 mL of water = 0.028 g.

If another 100 mL of water is added after equilibrium has been established, another 0.028 g of iodine would dissolve.

∴ mass of undissolved iodine – 0.30 – (0.028 x 2) = 0.30 – 0.056 = 0.244 g.

The concentration of iodine in the solution would simply be the solubility of iodine at equilibrium i. i.e., 0.28 gL-1.

Characteristics of equilibrium in physical processes: We have discussed the attainment of an equilibrium state in relation to four different physical processes. They all have certain characteristics in common.

  1. An equilibrium can be reached only in a dosed system, i.e., a system which neither gains matter from nor loses matter to the surroundings. For instance, in the case of a liquid—gas system, equilibrium is not reached if the vapour is allowed to escape from the vessel.
  2. The equilibrium reached is dynamic in nature, i.e., there is a balance between two opposing processes occurring at the same rate.
  3. At equilibrium, the measurable properties of the system become constant (under a given set of conditions). For instance, in the case of evaporation, the vapour pressure of the liquid is constant at a given temperature.
    • In solid-liquid equilibrium, the melting point (the temperature at which the two phases coexist provided no heat is exchanged) is constant for a particular pressure.
    • For the dissolution of a solid in a liquid, the solubility is constant at a given temperature. For the dissolution of a gas in a liquid, the concentration of the gas in the solution is constant for a particular pressure of the gas above the solution at a given temperature.
  4. At equilibrium, the concentrations of substances in the liquid and gaseous phase bear a constant ratio at a particular temperature. For the dissolution of C02 in water, for example, \(\frac{\mathrm{CO}_2(\mathrm{aq})}{\mathrm{CO}_2(\mathrm{~g})}\) is constant at a given temperature. This constant is called the equilibrium constant and its magnitude indicates the extent to which the process proceeds before equilibrium is reached. For example, the greater the value of \(\frac{\mathrm{CO}_2(\mathrm{aq})}{\mathrm{CO}_2(\mathrm{~g})}\) greater is the extent to which CO2 disiolves in water.

Equilibrium In Chemical Systems

We started this chapter with a discussion on whether a chemical reaction must always proceed to completion. There are some chemical reactions which proceed to completion, i.e., at equilibrium the concentration of the, reactants is negligible compared to the concentration of the products. (In reality, no reaction proceeds to completion in the true sense, but we need not concern ourselves with this now.) An example of this type of reaction is

⇒ \(\mathrm{BaCl}_2(\mathrm{aq})+\mathrm{Na}_2 \mathrm{SO}_4(\mathrm{aq}) \longrightarrow \mathrm{BaSO}_4(\mathrm{~s})+2 \mathrm{NaCl}(\mathrm{aq})\)

We could also call this reaction an irreversible reaction, since the reaction proceeds in one direction and the reverse process, i.e., the products reacting to form the reactants under the same conditions, does not take place. A few irreversible reactions are given below.

⇒ \(2 \mathrm{Na}+2 \mathrm{H}_2 \mathrm{O} \longrightarrow 2 \mathrm{NaOH}+\mathrm{H}_2\)

⇒ \(\mathrm{AgNO}_3(\mathrm{aq})+\mathrm{NaCl}(\mathrm{aq}) \longrightarrow \mathrm{NaNO}_3(\mathrm{aq})+\mathrm{AgCl}(\mathrm{s})\)

⇒ \(2 \mathrm{Mg}(\mathrm{s})+\mathrm{O}_2(\mathrm{~s}) \longrightarrow 2 \mathrm{MgO}(\mathrm{s})\)

Let us now consider a reaction which is reversible, i.e., a reaction in which the reactants react to form products under certain conditions and the products combine to form the reactants under the same conditions.

For instance, iron reacts with steam to form hydrogen and iron oxide, and heated iron oxide reacts with hydrogen to produce iron and steam.

⇒ \(3 \mathrm{Fe}+4 \mathrm{H}_2 \mathrm{O} \rightleftharpoons \mathrm{Fe}_3 \mathrm{O}_4+4 \mathrm{H}_2\)

If a reversible reaction is carried out in such a way that one of the products escapes, the reaction becomes irreversible, or proceeds in only one direction.

  • If, for instance, in the reaction between metallic iron and steam, the hydrogen formed is allowed to escape, the reverse reaction cannot take place and the process becomes virtually irreversible, or proceeds to completion.
  • If, on the other hand, the reaction is carried out in a closed vessel, from which hydrogen cannot escape, both forward and backward reactions take place simultaneously and there comes a time when an equilibrium is established.
  • At equilibrium, the reaction appears to have stopped, but what happens is that the forward and reverse reactions occur at the same rate.

For instance, when hydrogen and iodine are heated in a closed vessel at 717 K, they react to form hydrogen iodide. However, since the hydrogen iodide is not allowed to escape, the reverse reaction occurs. Hydrogen iodide decomposes to form hydrogen and iodine.

⇒ \(\underset{\text { colourless }}{\mathrm{H}_2(\mathrm{~g})+\mathrm{I}_2(\mathrm{~g})} \rightleftharpoons \underset{\text { viclet }}{2 \mathrm{HI}(\mathrm{g})}\)

Initially, the reaction mixture is deep violet because of the presence of iodine.

Basic Chemistry Class 11 Chapter 7 Equilibrium Chemical Equilibrium

  • The intensity of the colour decreases, as the reaction proceeds, and after a while, it becomes steady, indicating that an equilibrium has been reached.
  • At equilibrium, it appears that the reaction has stopped, despite the presence of the reactants.
  • But in reality, the rate at which hydrogen and iodine combine to form hydrogen iodide becomes equal to the rate at which hydrogen iodide decomposes.
  • When we talk about the time it takes for the colour of the
    reaction mixture that we have just been discussing to become
    steady, we are talking about a finite time interval.
  • However, to the naked eye, the change seems to be instantaneous. It is possible, though, to determine this time interval with modem instruments.
  • It is also possible to establish the dynamic nature of the equilibrium using radioactive reagents.

Characteristics of chemical equilibrium

1. Consider a general reversible reaction \(\mathrm{A}+\mathrm{B} \rightleftharpoons \mathrm{C}+\mathrm{D}\)

  • Initially, the concentrations of A and B are maximum and the concentrations of C and D are zero. As the reaction proceeds, the concentrations of A and B decrease and the concentrations of C and D increase.
  • With this, the rate of the forward reaction starts decreasing, while the rate of the backward reaction starts increasing, until the two rates become equal, and equilibrium is reached.
  • At this point the concentration of A and B, and C and D become constant.
  • That the concentrations of the reactants and the products become constant at equilibrium can be demonstrated by connecting a manometer to a closed vessel containing CaCO3.

If the vessel is heated, CaCO3 decomposes to yield CaO and CO2.

⇒ \(\mathrm{CaCO}_3(\mathrm{~s}) \rightleftharpoons \mathrm{CaO}(\mathrm{s})+\mathrm{CO}_2(\mathrm{~g})\)

To begin with, the pressure due to CO2 keeps increasing. Then the level of mercury in the manometer becomes steady, indicating that the pressure has become constant. This shows that the concentration of CO2 produced becomes constant (as those of CaCO3 and CaO) at equilibrium.

Basic Chemistry Class 11 Chapter 7 Equilibrium Decompostion Of CaCO3

2. In the reaction that we have just considered, the pressure exerted by CO2 becomes constant at equilibrium. In the reaction between hydrogen and iodine, the colour of the reaction mixture becomes constant.

In general, the observable properties of the system become constant at equilibrium. Actually, this follows from the fact that the concentrations of the reactants and products become constant at equilibrium.

3. A chemical equilibrium can be attained only in a closed system. This should be obvious. If the system is open, one or more of the products may escape, making the reverse reaction impossible.

4. The rates of the forward and backwards reactions become equal at equilibrium. We have discussed this point at length already.

If this were not so, the reaction between hydrogen and iodine, for instance, would not appear to come to a standstill, despite the presence of the reactants.

The dynamic nature of chemical equilibrium can be demonstrated in the synthesis of ammonia by the Haber process.

⇒ \(\mathrm{N}_2(\mathrm{~g})+3 \mathrm{H}_2(\mathrm{~g}) \rightleftharpoons 2 \mathrm{NH}_3(\mathrm{~g})\)

Known amounts of hydrogen and nitrogen are taken at high temperature and pressure, which react to form ammonia.

  • The concentrations of all the three substances respectively are determined at regular intervals and a graph of the molar concentrations against time is plotted.
  • The graph shows that initially the concentration of the reactants, i.c„ hydrogen and nitrogen decreases whereas that of the product ammonia increases.
  • That equilibrium is reached is depicted in the graph when, after a certain time, the composition of the reaction mixture is constant.
  • In order to prove the dynamic nature of equilibrium, the experiment is performed again in the same set of conditions except for the use of deuterium in place of hydrogen.
  • The reaction mixture in this case, at equilibrium, has the same composition as that in the prior one except for the presence of deuterium in place of hydrogen (i.e., D2, N2 and ND3).
  • Once the equilibrium is attained in both cases, the two reaction mixtures are mixed together and left. After some time the concentration of ammonia in this new mixture is found to be the same as before.
  • However, mass spectrometer analysis revealed that all forms containing hydrogen as well as deuterium are present—NH3, NH2D, NHD2, ND3, H2, HD, D2—in the reaction mixture.
  • This is only possible when the bonds are continuously breaking and being made at equilibrium or the forward and reverse reactions are taking place.

Had the reaction stopped then there would have been no mixing of the isotopes. This proves that the equilibrium is dynamic in nature.

Basic Chemistry Class 11 Chapter 7 Equilibrium Plot Of Molar Concentartions Against Time For Components Of Reaction Mixture In Haber Process

5. The equilibrium can be approached from either direction. For instance, in the case of \(\mathrm{Fe}_3 \mathrm{O}_4+4 \mathrm{H}_2 \rightleftharpoons 3 \mathrm{Fe}+4 \mathrm{H}_2 \mathrm{O}\), it would make no difference whether we started with iron oxide and hydrogen, or metallic iron and steam.

  • The fact that equilibrium can be approached from either direction can be proved by performing a simple experiment. Dinitrogen tetroxide (N2O4) is a colourless gas and nitrogen dioxide (NO2) is reddish brown in colour.
  • Dinitrogen tetroxide decomposes almost completely to nitrogen dioxide at 373 K. N2O4 is stable at 273 K and exists almost entirely as pure N2O4, which is colourless.

⇒ \(\underset{\text { collourless }}{\mathrm{N}_2 \mathrm{O}_4(\mathrm{~g})} \rightleftharpoons \underset{\text { reddish brown }}{2 \mathrm{NO}_2(\mathrm{~g})}\)

If you take two identical flasks A and B, fill them with nitrogen dioxide and keep them at room temperature, both will actually contain a pale brown mixture of N2O4 and NO2.

  • Now place flask A in a bath maintained at 273 K and flask B in a vessel of boiling water. The gas in flask A will turn colourless, indicating that it is mostly N2O4, while the colour of the gas in flask B will deepen to reddish brown, showing that it is mostly NO2.
  • If you then transfer both the flasks to a bath maintained at room temperature (298 K), the gas in flask A will turn pale brown and the reddish brown colour of flask B will fade, until both appear the same colour.
  • The fact that the gaseous mixtures in both flasks attain the same colour indicates that, at equilibrium, both flasks contain a mixture of N2O4 and NO2 in the same proportion, which means that equilibrium can be approached from either direction.

The fact is also demonstrated by Figure, which graphically represents the reaction \(\mathrm{H}_2+\mathrm{I}_2 \rightleftharpoons 2 \mathrm{HI}\) we have considered earlier in the chapter.

Basic Chemistry Class 11 Chapter 7 Equilibrium Experiment To Show That Equilibrium Can Be Approcahced From Either Direction

Whether we start the reaction from H2 and I2 or from HI, the composition of the reaction mixture at equilibrium is constant. The figure represents the forward and the backward reaction in the same graph.

  • When we start the reaction with an equimolar mixture of H2 and I2, as you can see in Figure, the concentration of these two decreases (curve a) while that of HI increases (curve b). Both of them finally reach a stable concentration at equilibrium.
  • When we move from right to left in the graph, we are actually starting with HI and its concentration decreases with time (curve c) while that of H2 and I2 increases (curve d), both of them stabilising at the equilibrium point.

Basic Chemistry Class 11 Chapter 7 Equilibrium Chemical Equilibnrium In The Reaction Can Be Attained From Either Direction

6. You know that a catalyst increases the rate of a reaction. In the case of a reversible reaction, a catalyst increases the rates of the forward and reverse reactions to the same extent.

Hence it does not alter the state of equilibrium. That is to say, the concentrations of the reactants and products at equilibrium are the same with or without the catalyst. All that happens is that the state of equilibrium is attained faster.

7. To repeat what you already know, at equilibrium, ΔG = 0.

Law Of Mass Action

We have been discussing the fact that at equilibrium, the rates of the forward and reverse reactions become equal. But can one determine these rates? In 1863 two Norwegian chemists, C. M. Guldberg and P. Waage, came up with a law on the basis of experimental observations.

  • This law, known as the law of mass action, relates the rate of chemical reaction with the concentrations of the reactants. It says that the rate at which a chemical reaction occurs at a given temperature is proportional to the product of the active masses of the reactants.
  • You must remember that Guldberg and Waage came up with this law in connection with ideal gases and that the law is strictly correct only for ideal gases.

The active mass of a reactant is nothing but its molar concentration, or the number of moles dissolved per litre of the solution. For example, suppose a solution of HCl has x g of HCl dissolved in y L of solution. Then the concentration of the solution = \(\frac{x}{y} \mathrm{gL}^{-1}\)

= \(\frac{x}{36.5 \times y} \mathrm{~mol} \mathrm{~L}^{-1}\)(molar mass of HCl = 36.5)

Thus the active mass of the HCl solution, represented as [HCl], is x/36.5 y M, where M stands for molar concentration or mol L-1.

Let us consider a reaction between two reactants A and B.

A + B → products

According to the law of mass action, the rate of reaction ∝ [A] [B] or rate = k[A][B],

where [A] and [B] are the concentrations of the reactants A and B respectively, and A is a proportionality constant, called the velocity constant or rate constant.

Now consider a reaction 3C + 4D → products

you could think of this reaction as C + C + C + D + D + D + D → products then rate = k[C][C][C][D][D][D][D] = k[C]3[D]4.

In general, for any reaction aA + bB → products

∴ rate of reaction = k[A]a [B]b.

Thus, the law of mass action can also be expressed as follows.

The rate of a reaction is proportional to the product of the concentrations of the reactants, each raised to the power equal to its coefficient (number of moles of the species) in the balanced chemical equation representing the reaction.

Law Of Chemical Equilibrium

This law follows from the law of mass action. Consider the following general reversible reaction.

aA + bB \(\rightleftharpoons\)cC + dD

If we apply the law of mass action to this reaction at equilibrium, rate of forward reaction = \(k[\mathrm{~A}]_{\mathrm{eq}}^a[\mathrm{~B}]_{\mathrm{eq}}^b\),

where k is the rate constant for the forward reaction and [A]eq and [B]eq are the molar concentrations of the reactants A and B at equilibrium.

Rate of reverse reaction = \(k^{\prime}[C]_{\mathrm{eq}}^c[\mathrm{D}]_{\mathrm{eq}}^d\),

where k’ is the rate constant for the reverse reaction and [C]eq and [D]eq are the molar concentrations of the products C and D at equilibrium.

At equilibrium, rate of forward reaction = rate of reverse reaction or \(k[\mathrm{~A}]_{\mathrm{eq}}^a[\mathrm{~B}]_{\mathrm{eq}}^b=k^{\prime}[\mathrm{C}]_{\mathrm{eq}}^c[\mathrm{D}]_{\mathrm{eq}}^d\)

or, \(\frac{k}{k^{\prime}}=\frac{[\mathrm{C}]_{\mathrm{eq}}^c[\mathrm{D}]_{\mathrm{eq}}^d}{[\mathrm{~A}]_{\mathrm{eq}}^a[\mathrm{~B}]_{\mathrm{eq}}^b}=K_c\)

Generally the subscript ‘eq’ is used to denote concentration at equilibrium. The subscript ‘c’ indicates that Kc is expressed in concentrations of mol L-1.

Kc must be a constant at a given temperature since k and Ak’ are constants for a given temperature. Guldberg and Waage called it the equilibrium constant.

Thus, at equilibrium, the product of the molar concentrations of the reaction products, each raised to the power equal to its stoichiometric coefficient in the balanced chemical equation, divided by the product of the molar concentrations of the reactants, each raised to the power equal to its stoichiometric coefficient, is constant at a given temperature. This is referred to as the law of chemical equilibrium.

The equilibrium constant for the reaction \(4 \mathrm{NH}_3(\mathrm{~g})+5 \mathrm{O}_2(\mathrm{~g}) \rightleftharpoons 4 \mathrm{NO}(\mathrm{g})+6 \mathrm{H}_2 \mathrm{O}(\mathrm{g}) \text { is } K_{\mathrm{c}}=\frac{\left[\mathrm{NO}^4\left[\mathrm{H}_2 \mathrm{O}\right]^6\right.}{\left[\mathrm{NH}_3\right]^4\left[\mathrm{O}_2\right]^5}\)

The symbols used to denote phases in a chemical reaction are ignored while writing the expression for the equilibrium constant.

Reaction quotient: For any reversible reaction at any stage other than equilibrium, the ratio of the molar concentrations of the products to that of the reactants, where each concentration term is raised to the power equal to the stoichiometric coefficient of the substance concerned, is called the reaction quotient, Q.

It is symbolised as Qc. while molar concentrations are considered and Qp while partial pressures are considered for any reaction.

For a general reaction \(a\mathrm{~A}+b \mathrm{~B} \rightleftharpoons c \mathrm{C}+d \mathrm{D}\)

which is not at equilibrium  \(Q_c=\frac{[\mathrm{C}]^c[\mathrm{D}]^d}{[\mathrm{~A}]^a[\mathrm{~B}]^b}\).

  1. If Qc > Kc, the value of Qc will tend to decrease to reach the value of Kc (towards equilibrium), and the reaction will proceed in the reverse direction.
  2. If Qc < Kc, it will tend to increase, and the reaction will proceed in the forward direction.
  3. If Qc = Kc, the reaction is at equilibrium.

So, if you know the relative concentrations of reactants and products at a particular stage of a reaction, you can predict the direction in which the reaction will proceed.

Also, if you know the initial concentrations of the reactants and the value of Kc, you can predict the composition of the reaction system at equilibrium.

Now consider the following reaction. \(\mathrm{CO}(\mathrm{g})+\mathrm{H}_2 \mathrm{O}(\mathrm{g}) \rightleftharpoons \mathrm{CO}_2(\mathrm{~g})+\mathrm{H}_2(\mathrm{~g})\)

Here \(Q_c=\frac{\left[\mathrm{CO}_2\right]\left[\mathrm{H}_2\right]}{[\mathrm{CO}]\left[\mathrm{H}_2 \mathrm{O}\right]}\) and Kc = 0.64 at 800°C. The initial conditions for a reaction can vary.

Here we have discussed three possibilities for the given reaction. The values taken are from different experiments. There is no need for you to memorise these values, the discussion will help you learn how to predict the direction of a reaction.

1. Only the reactants are present, say. [CO] = 0.0243 M and [H2O] = 0.0243 M.

Since the concentration of products is zero, the reaction quotient Qc = 0, i.e., Qc < Kc.

The reaction thus proceeds in the forward direction till Qc reaches a value equal to Kc.

At equilibrium, the concentrations are [CO]eq = 0.0135 M,[H2O]eq = 0.0135 M,[CO2]eq = 0.0108 M,[H2]eq = 0.0108 M.

∴ \(K_c=\frac{0.0108 \times 0.0108}{0.0135 \times 0.0135}=0.640\)

2. Only the products arc present, say. [CO2] = 0.0468 M,[H2] = 0.0468 M.

Qc >Kc and hence the reaction proceeds in the reverse direction, forming CO and H2O.

At equilibrium the concentrations are [CO]eq = 0.0260 M,[H2O]eq = 0.0260 M,[CO2]eq = 0.0208 M,[H2]eq = 0.0208 M.

∴ \(K_2=\frac{(0.0208)(0.0208)}{(0.0260)(0.0260)}\)=0.640

3. The reactants and products are present in the following concentrations.

[CO]= 0.0094 M.[HO] =0.055 M,[CO2] = 0.0005 M,[H2]= 0.0046 M.

then, \(Q_c=\frac{(0.0006)(0.0046)}{(0.0094)(0.0065)}=0.044\)

As Qc. <Kc, the reaction will proceeds in the forward direction and as expected, the concentrations at equilibrium are found to be [CO]eq = 0.0074 M; [H2O]eq = 0.0035 M; [CO2]eq =0.0025 M and [H2]eq = 0.0066 M resulting in a Kc value of 0.64.

Types of equilibrium: A chemical equilibrium is called homogeneous if all the reactants and products are present in the same phase. Some examples of homogeneous equilibrium follow.

⇒ \(\mathrm{N}_2(\mathrm{~g})+3 \mathrm{H}_2(\mathrm{~g})\rightleftharpoons 2 \mathrm{NH}_3(\mathrm{~g})\)

⇒ \(\mathrm{CO}(\mathrm{g})+\mathrm{H}_2 \mathrm{O}(\mathrm{g}) \rightleftharpoons \mathrm{CO}_2(\mathrm{~g})+\mathrm{H}_2\)

⇒ \(\mathrm{CH}_3 \mathrm{COOH}(\mathrm{aq})+\mathrm{C}_2 \mathrm{H}_5 \mathrm{OH}(\mathrm{aq}) \rightleftharpoons \mathrm{CH}_3 \mathrm{COOC}_2 \mathrm{H}_5(\mathrm{aq})+\mathrm{H}_2 \mathrm{O}(\mathrm{l})\)

In a heterogeneous, chemical equilibrium, on the other hand, the reactants and products are not all in the same phases A few examples follow.

⇒ \(\mathrm{CaCO}_3(\mathrm{~s}) \rightleftharpoons \mathrm{CaO}(\mathrm{s})+\mathrm{CO}_2(\mathrm{~g})\)

⇒ \(\mathrm{C}(\mathrm{s})+\mathrm{H}_2 \mathrm{O}(\mathrm{g}) \rightleftharpoons \mathrm{CO}(\mathrm{g})+\mathrm{H}_2(\mathrm{~g})\)

⇒ \(3 \mathrm{Fe}(\mathrm{s})+4 \mathrm{H}_2 \mathrm{O}(\mathrm{g}) \rightleftharpoons \mathrm{Fe}_3 \mathrm{O}_4(\mathrm{~s})+4 \mathrm{H}_2(\mathrm{~g})\)

∴ \(\mathrm{CrO}_4^{2-}(\mathrm{aq})+\mathrm{Pb}^{2+}(\mathrm{aq}) \rightleftharpoons \mathrm{PbCrO}_4(\mathrm{~s})\)

Gas-phase reactions: When all fee reactants and products of a reaction are gaseous, the reaction is called a gas-phase reaction and it is a case of homogeneous equilibrium.

In the case of such reactions, the equilibrium constant can be expressed either in terms of molar concentrations or partial pressures.

When expressed in terms of molar concentrations it is denoted by Kc and if it is expressed in terms of partial pressures, Kp is used. If A, B, X and Y are gases in the reaction the equilibrium constant can be expressed as follows.

⇒ \(a \mathrm{~A}+b \mathrm{~B} \rightleftharpoons x \mathrm{X}+y \mathrm{Y}\)

∴ \(K_p=\frac{\left(p_{\mathrm{X}}\right)^x\left(p_{\mathrm{Y}}\right)^y}{\left(p_{\mathrm{A}}\right)^a\left(p_{\mathrm{B}}\right)^{\mathrm{b}}}\)

In this expression, pA, pB, px and py are the partial pressures of A, B, X and Y, expressed in atmospheres or pascals.

At constant temperature, the pressure of a gas is proportional to its concentration.

In terms of concentrations, the equilibrium constant can be expressed as

∴ \(K_c=\frac{[X]^z[Y]^y}{[A]^a[B]^b} \quad \text { or } \quad \frac{\left(C_X\right)^x\left(C_Y\right)^y}{\left(C_A\right)^a\left(C_B\right)^b} \text {, }\)

where CA, CB, Cx and Cy are the molar concentrations of A, B, X and Y respectively.

If A, B, X and Y are ideal gases, pV = nRT

or \(p=\frac{n}{V} R T\)

or p =n/VRT = CRT [n/V = number of moles per litre or mol/dm3]

Then \(p_{\mathrm{A}}\)=\(\mathrm{C}_{\mathrm{A}} R T\) and \(p_{\mathrm{B}}\)=\(\mathrm{C}_{\mathrm{B}} R T\).

⇒ \(p_{\mathrm{X}}\)=\(C_{\mathrm{X}} R T\) and \(p_{\mathrm{Y}}\)=\(\mathrm{C}_{\mathrm{Y}} R T\).

∴ \(K_p=\frac{\left(C_X R T\right)^x\left(C_Y R T\right)^y}{\left(C_A R T\right)^a\left(C_B R T\right)^b}=\frac{\left(C_X\right)^x\left(C_Y\right)^y}{\left(C_{\mathrm{A}}\right)^a\left(C_B\right)^b} R T^{(x+y)-(a+b)}\).

But \(K_c=\frac{\left(C_X\right)^x\left(C_Y\right)^y}{\left(C_{\mathrm{A}}\right)^a\left(C_B\right)^b}\).

∴ \(K_p=K_c R T^{\Delta n}\),

where Δn = (x + y) – (a + b) = no. of moles of products – no. of moles of reactants in the balanced chemical equation.

Consider the reaction \(\mathrm{H}_2(\mathrm{~g})+\mathrm{I}_2(\mathrm{~g}) \rightleftharpoons 2 \mathrm{HI}(\mathrm{g})\) at equilibrium. If the equilibrium constant for the reaction is Kc then Kp would be Kp = KcRTΔn,

where Δn = no. of moles of HI- no. of moles of H2 and I2 = 2- 2 or 0.

Therefore, Kp =Kc for the reaction at equilibrium.

However, this is not always the case. Consider another example, \(\mathrm{N}_2(\mathrm{~g})+3 \mathrm{H}_2(\mathrm{~g}) \rightleftharpoons 2 \mathrm{NH}_3(\mathrm{~g})\) at equilibrium.

Here the number of moles of reactants is more than that of the product or Δn = 2 – 4 or -2.

Therefore, the equilibrium constant in terms of partial pressure Kp for the above reaction would be Kp = Kc(RT)-2.

If the concentration is in mol L-1 and p is in bar then R = 0.0831 L bar K-1 mol-1.

Nature of equilibrium constant

1. The equilibrium constant, as we have defined it so far, is the ratio of the product of the molar concentrations of the reaction products to that of the reactants, with each concentration term raised to the power equal to the stoichiometric coefficient of the substance in the balanced chemical equation. You may argue that if this is so, K should have units.

  • Though the equilibrium constant was initially derived in terms of concentrations, strictly speaking, concentrations or partial pressures can be used only for ideal gases. However, nowadays, equilibrium constants are expressed in dimensionless quantities.
  • This is possible only when the quantities (concentration and partial pressures) are measured with respect to a corresponding standard state.
  • The standard state for a pure gas is taken to be 1 bar and for solute the standard state (C0) is 1 molar solution. Thus a pressure of 4 bar in terms of standard state is 4, a number without a unit.
  • Therefore the numerical value of the equilibrium constant depends on the standard state chosen and both Kp and Kc are dimensionless quantities.

2. The equilibrium constant has a definite value for a particular reaction at a given temperature. This value does not depend on the initial concentrations of the reactants.

If the reaction is reversed, the tire value of K becomes the inverse of the value of the forward process. For instance, suppose the value of the equilibrium constant for the combination of hydrogen and iodine at 720 K is Kc.

⇒ \(\mathrm{H}_2(\mathrm{~g})+\mathrm{I}_2(\mathrm{~g}) \rightleftharpoons 2 \mathrm{HI}(\mathrm{g}) ; \quad K_c=\frac{[\mathrm{HI}]^2}{\left[\mathrm{H}_2\right]\left[\mathrm{I}_2\right]} \text {. }\)

Then the value of the equilibrium constant, say K’c, for the decomposition of hydrogen iodide at the same temperature would be 1/Kc.

⇒ \(2 \mathrm{HI}(\mathrm{g}) \rightleftharpoons \mathrm{H}_2(\mathrm{~g})+\mathrm{I}_2(\mathrm{~g}) ; \quad K_c^{\prime}=\frac{1}{K_c}=\frac{\left[\mathrm{H}_2\right]\left[\mathrm{I}_2\right]}{[\mathrm{HI}]^2}\).

3. It is all very well to say that the value of the equilibrium constant remains the same for a particular reaction at a given temperature. But a reaction may be expressed in more ways than one. For instance, the reaction between hydrogen and iodine can be expressed in the following ways.

⇒ \(\mathrm{H}_2(\mathrm{~g})+\mathrm{I}_2(\mathrm{~g}) \rightleftharpoons 2 \mathrm{HI}(\mathrm{g})\)….(1)

⇒ \(\frac{1}{2} \mathrm{H}_2(\mathrm{~g})+\frac{1}{2} \mathrm{I}_2(\mathrm{~g}) \rightleftharpoons \mathrm{HI}(\mathrm{g})\)…(2)

⇒ \(n \mathrm{H}_2(\mathrm{~g})+n \mathrm{I}_2(\mathrm{~g}) \rightleftharpoons 2 n \mathrm{HI}(\mathrm{g})\) …..(3)

If the equilibrium constant for (1) is \(K_c=\frac{[\mathrm{HI}]^2}{\left[\mathrm{H}_2\right]\left[\mathrm{I}_2\right]}\)

then the equilibrium constant for (2) is \(K_c^{\prime}=\frac{[\mathrm{HI}]}{\left[\mathrm{H}_2\right]^{1 / 2}\left[\mathrm{I}_2\right]^{1 / 2}}=\sqrt{K_c} \text { or } K_c^{1 / 2}\)

and the equilibrium for  (3) is \(K_c^n=\frac{[\mathrm{HI}]^{2 n}}{\left[\mathrm{H}_2\right]^n\left[\mathrm{I}_2\right]^n}=K_c^n\).

Thus, the value of Kc depends on the coefficients of the reactants and products in the equation being considered.

4. If an equation is written in two steps and the values of the equilibrium constants for the two steps are K1 and K2 then the value of the equilibrium constant for the equation is the product of and K2. For example, the reaction

⇒ \(\mathrm{N}_2+2 \mathrm{O}_2 \rightleftharpoons 2 \mathrm{NO}_2 ; \quad K_c=\left[\mathrm{NO}_2\right]^2 /\left[\mathrm{N}_2\right]\left[\mathrm{O}_2\right]^2\)

occurs in two steps.

⇒ \(\mathrm{N}_2+\mathrm{O}_2 \rightleftharpoons 2 \mathrm{NO} ; \quad K_1=\left[\mathrm{NO}^2 /\left[\mathrm{N}_2\right]\left[\mathrm{O}_2\right]\right.\)

⇒ \(2 \mathrm{NO}+\mathrm{O}_2 \rightleftharpoons 2 \mathrm{NO}_2 ; \quad K_2=\left[\mathrm{NO}_2\right]^2 /\left[\mathrm{NO}^2\left[\mathrm{O}_2\right]\right.\)

∴ \(K_1K_2=\frac{[\mathrm{NO}]^2}{\left[\mathrm{~N}_2\right]\left[\mathrm{O}_2\right]} \frac{\left[\mathrm{NO}_2\right]^2}{\left[\mathrm{NO}^2\left[\mathrm{O}_2\right]\right.}=\frac{\left[\mathrm{NO}_2\right]^2}{\left[\mathrm{~N}_2\right]\left[\mathrm{O}_2\right]^2}=K_c\)

5. The value of equilibrium constant does not change if a catalyst is present.

6. The magnitude of Kc helps to predict the extent to which the reactants will be converted into the products, or the equilibrium composition of the reaction system.

  • As products appear in the numerator of the equilibrium constant expression, and reactants are in the denominator, a large value of Kc or Kp indicates that the ratio of products to reactants is very large.
  • In other words, the reaction proceeds nearly to completion. A very low value of Kc indicates that the reaction cannot proceed much in the forward direction.

In other words the value of Kc tells us whether the particular reaction can actually take place. Let us consider an example.

⇒ \(\mathrm{Fe}^{3+}(\mathrm{aq})+\mathrm{SCN}^{-}(\mathrm{aq}) \rightleftharpoons \mathrm{FeSCN}^{2+}(\mathrm{aq})\)

At 298 K, the equilibrium constant for the forward reaction is 138, which shows that the forward reaction is favoured or can occur. For the reverse reaction the equilibrium constant is 1/138, which shows that this reaction does not proceed much to the left.

Similarly, for the reaction, \(\mathrm{H}_2(\mathrm{~g})+\mathrm{Cl}_2(\mathrm{~g}) \rightleftharpoons 2 \mathrm{HCl} \text { at } 300 \mathrm{~K}\) the equilibrium constant is 4.0 x 1031. This shows that the reaction goes virtually to completion.

The equilibrium constant for the reverse reaction \(2 \mathrm{HCl}(\mathrm{g}) \rightleftharpoons \mathrm{H}_2(\mathrm{~g})+\mathrm{Cl}_2(\mathrm{~g}) \text { is } \frac{1}{4.0 \times 10^{31}}=25 \times 10^{-30}\) indicating that the reverse reaction is less favoured. While a large value of Kp or Kc favours the formation of products, a very small value favours the reverse reaction or the formation of reactants.

Consider the reaction \(\mathrm{N}_2(\mathrm{~g})+\mathrm{O}_2(\mathrm{~g}) \rightleftharpoons 2 \mathrm{NO}(\mathrm{g}) . \mathrm{K}_c \text { is } 4.8 \times 10^{-31}\) at 298 K for this reaction. Here, at equilibrium, the ratio of concentration of products to reactants is small, indicating that the reverse reaction is favoured.

Examples of reactions having intermediate values of equilibrium constants are

⇒ \(\mathrm{H}_2(\mathrm{~g})+\mathrm{I}_2(\mathrm{~g}) \rightleftharpoons 2 \mathrm{HI}(\mathrm{g}) K_c=57 \text { at } 700 \mathrm{~K}\)

⇒ \(\text { and } \mathrm{N}_2 \mathrm{O}_4(\mathrm{~g})  \rightleftharpoons 2 \mathrm{NO}_2(\mathrm{~g}) K_c=4.64 \times 10^{-3} \text { at } 298 \mathrm{~K}\).

In both cases, there are appreciable amounts of reactants and products. This may be verified for the first reaction by substituting the concentrations of H2 and I2 (say 0.01 M) in the equilibrium constant expression

⇒ \(K_c=\frac{[\mathrm{HI}]^2}{\left[\mathrm{H}_2\right]\left[\mathrm{I}_2\right]} \cdot\)

⇒ \({[\mathrm{HI}]^2 } =K_c \cdot\left[\mathrm{H}_2\right]\left[\mathrm{I}_2\right]\)

∴ \({[\mathrm{HI}] } =\sqrt{K_c \cdot\left[\mathrm{H}_2\right]\left[\mathrm{I}_2\right]}=\sqrt{57 \times 0.01 \times 0.01}=0.075 \mathrm{M}\).

It is worthwhile to mention here that the magnitude of an equilibrium constant of a reaction does not tell us anything about the rate of that reaction, i.e., how fast or how slowly the reaction will take place. It is quite possible that the equilibrium constant of a reaction may suggest completion but the rate is too low for all practical purposes.

We can make a generalisation about the extent to which a reaction proceeds based on the magnitude of Kc, as shown in Table.

Basic Chemistry Class 11 Chapter 7 Equilibrium Values Of Kp Or Kc Predict The Extent To Which A Reaction Proceeds

Conventions regarding K

1. In the case of a homogeneous equilibrium involving products and reactants in the gaseous phase for example, \(\mathrm{H}_2(\mathrm{~g})+\mathrm{I}_2(\mathrm{~g}) \rightleftharpoons 2 \mathrm{HI}(\mathrm{g})\)

obtaining the expression for Kc or Kp is simple enough.

∴ \(K_{\mathrm{c}}=\frac{[\mathrm{HI}(\mathrm{g})]^2}{\left[\mathrm{H}_2(\mathrm{~g})\right]\left[\mathrm{I}_2(\mathrm{~g})\right]} \text { and } K_p=\frac{p_{\mathrm{HI}}^2}{p_{\mathrm{H}_2} p_{\mathrm{I}_2}} \text {. }\)

2. If one of the products or reactants involved in a heterogeneous equilibrium is a solid or a liquid its concentration is taken to be unity, by convention.

The molar concentration of a pure solid or liquid is coastant. In other words for any amount of substance ‘X’, [X(s)] and [X(l)] are constant. But [X(g)] and [X(aq)] will vary with the volume. Consider the following equilibrium.

⇒ \(\mathrm{CaCO}_3(\mathrm{~s}) \rightleftharpoons \mathrm{CaO}(\mathrm{s})+\mathrm{CO}_2(\mathrm{~g})\)

⇒ \(K_{\mathrm{c}}=\frac{[\mathrm{CaO}(\mathrm{s})]\left[\mathrm{CO}_2(\mathrm{~g})\right]}{\left[\mathrm{CaCO}_3(\mathrm{~s})\right]}=\frac{1 \times\left[\mathrm{CO}_2(\mathrm{~g})\right]}{1}=\left[\mathrm{CO}_2(\mathrm{~g})\right]\)

or \(K_p=p_{\mathrm{CO}_2}\)

This fits with the fact that the pressure of CO2 becomes constant when equilibrium is reached in the decomposition of CaCO3 in a closed vessel.

It has been found experimentally that the pressure of CO2 is 2.5 x 104 Pa when the reaction for the decomposition of CaCO3 reaches equilibrium at 1073 K. Therefore the equilibrium constant at 1073 K for the above reaction is

⇒ \(K_p=p_{\mathrm{CO}_2}=\frac{2.5 \times 10^4 \mathrm{~Pa}}{10^5 \mathrm{~Pa}}\) = 0.25

(since 1 bar =105 Pa and while calculating the value of Kp, pressure should be expressed in bar as the standard state of a substance is its pure state at exactly 1 bar.)

3. Consider an equilibrium in an aqueous medium.

⇒ \(\mathrm{NH}_3(\mathrm{aq})+\mathrm{H}_2 \mathrm{O}(\mathrm{l}) \rightleftharpoons \mathrm{NH}_4^{+}(\mathrm{aq})+\mathrm{OH}^{-}(\mathrm{aq})\)

In this case H2O is the solvent and is present in a large quantity. Consequently, its concentration does not change much during the reaction and is taken to be constant (unity, by convention). In general, the concentration of a solvent is taken to be 1.

Thus, \(K_c=\frac{\left[\mathrm{NH}_4^{+}(\mathrm{aq})\right]\left[\mathrm{OH}^{-}(\mathrm{aq})\right]}{\left[\mathrm{NH}_3(\mathrm{aq})\right] \times 1}\).

4. In the light of what we have just discussed, let us reconsider the equilibrium between a liquid and its vapour in a closed system.

⇒ \(\mathrm{H}_2 \mathrm{O}(\mathrm{l}) \rightleftharpoons \mathrm{H}_2 \mathrm{O}(\mathrm{g})\)

⇒ \(K_c=\frac{\left[\mathrm{H}_2 \mathrm{O}(\mathrm{g})\right]}{\left[\mathrm{H}_2 \mathrm{O}(\mathrm{l})\right]}=\left[\mathrm{H}_2 \mathrm{O}(\mathrm{g})\right]\)

or \(K_p=p_{\mathrm{H}_2 \mathrm{O}}\).

Now do you see why the vapour pressure of a liquid is constant at a particular temperature and why it does not depend on the amount of liquid present.

5. Remember that in a chemical equilibrium the concentration of a liquid is taken as 1 only if the liquid is present as a solvent, or in a large quantity. Consider the following reaction.

⇒ \(\mathrm{CH}_3 \mathrm{COOH}(\mathrm{l})+\mathrm{C}_2 \mathrm{H}_5 \mathrm{OH}(\mathrm{l}) \rightleftharpoons \mathrm{CH}_3 \mathrm{COOC}_2 \mathrm{H}_5(\mathrm{l})+\mathrm{H}_2 \mathrm{O}(\mathrm{l})\)

Here \(K=\frac{\left[\mathrm{CH}_3 \mathrm{COOC}_2 \mathrm{H}_5(\mathrm{l})\right]\left[\mathrm{H}_2 \mathrm{O}(\mathrm{l})\right]}{\left[\mathrm{CH}_3 \mathrm{COOH}(\mathrm{l})\right]\left[\mathrm{C}_2 \mathrm{H}_5 \mathrm{OH}(\mathrm{l})\right]}\)

The concentration of H2O cannot be taken as 1 in this case.

Example: A reaction between hydrogen and iodine to produce hydrogen iodide at 675 K is carried out in a closed flask of volume 4 L. At equilibrium, the reaction system contains 0.6 mol of hydrogen, 0.6 mol of iodine and 2.4 mol ofhydrogen iodide. Calculate the equilibrium constant.
Solution:

For the reaction \(\mathrm{H}_2(\mathrm{~g})+\mathrm{I}_2(\mathrm{~g}) \rightleftharpoons 2 \mathrm{HI}(\mathrm{g})\)

∴ \(K_{\mathrm{c}}=\frac{[\mathrm{HI}]^2}{\left[\mathrm{H}_2\right]\left[\mathrm{I}_2\right]}\)

Given that,

[HI] = 2.4/4 = 0.6 mol L-1.

[H2] = 0.6/4 = 0.15 mol L-1.

[I2] = 0.6/4 = 0.15 mol L-1.

Substituting these values in the expression for Kc, \(K_c=\frac{(0.6)^2}{0.15 \times 0.15}=16\)

The equilibrium constant, the reaction quotient and Gibbs energy: As already stated, the value of the equilibrium constant for a reaction does not depend on the rate of that reaction or vice versa.

  • However, it is directly related to the thermodynamics of the reaction and, in particular, to the change in free energy ΔG. You already know from the previous chapter that if ΔG is negative for a reaction then the reaction is spontaneous and should proceed in the forward direction.
  • If, on the other hand, ΔG is positive, the reaction is considered nonspontaneous. Then the reverse reaction will have a negative ΔG value and will proceed, converting the products of the forward reaction into reactants.
  • Either way, a reaction should proceed in the spontaneous direction until it achieves equilibrium. At this point, there is no free energy left to drive the reaction ΔG becomes zero and the reaction attains equilibrium.

This thermodynamic view of equilibrium is expressed mathematically as \(\Delta G=\Delta G^{\ominus}+R T\) ln Q

When equilibrium is achieved ΔG = 0, Q = K and this equation becomes \(\Delta G =\Delta G^{\ominus}+R T \ln K=0\)

or \(\Delta G^{\ominus} =-R T \ln K\).

Rearranging the above equation, we get

ln \(K=-\frac{\Delta G^{\ominus}}{R T}\)

or \(2.303 \log K =-\frac{\Delta G^{\ominus}}{R T}\)

or \(\log K =-\frac{\Delta G^{\ominus}}{2303 R T}\).

Taking antilog of both sides K = \(e^{-\Delta G^{\ominus} / R T}\)

or \(K=\mathrm{antilog}\left(-\frac{\Delta G^{\ominus}}{2.303 R T}\right)\)

This equation provides an interpretation of the spontaneity of a reaction.

If \(\Delta G^{\ominus}<0 \text { then } \frac{-\Delta G^{\ominus}}{R T}\) is positive, and \(e^{-\Delta C^{\ominus} / R T}>1 \text {, i.e., } K>1\). Such a reaction is spontaneous and proceeds in the forward direction so that the products are present predominantly.

Similarly, a nonspontaneous reaction is one for which \(\Delta G^\ominus>0\) and K < 1 Such a reaction proceeds in the forward direction to such a small extent that very small amounts of products are formed.

Example 1. Consider the reaction \(2 \mathrm{SO}_2(\mathrm{~g})+\mathrm{O}_2(\mathrm{~g}) \rightleftharpoons 2 \mathrm{SO}_3(\mathrm{~g})\). Amixture of SO2, SO3 and O2 at equilibrium at 1000 K has [SO2] = 3.8X10-3 M, [SO3] = 4.13X10-3 M, and [O2] = 4.3×10-3 M. Calculate Kc and Kp of the reaction.
Solution:

⇒ \(K_c=\frac{\left[\mathrm{SO}_3\right]^2}{\left[\mathrm{SO}_2\right]^2\left[\mathrm{O}_2\right]}\)

Substituting the values in the expression, we get \(K_c=\frac{\left(4.13 \times 10^{-3}\right)^2}{\left(3.8 \times 10^{-3}\right)^2\left(4.3 \times 10^{-3}\right)}\)

(Note: The concentration must always be in ‘M’ or mol L-1 for calculating Kc.]

⇒ Kc = 275.

To calculate Kp we must know Δn.

Δn = no. of moles of products- no. of moles of reactants. (considering only gaseous products and reactants)

Δn = 2- (2 + 1) = 2- 3 = -1

But Kp = Kc(RT)Δn

= Kc(RT)-1

= \(\frac{K_c}{R T}\)

∴ \(K_p=\frac{275}{(0.08312)(1000)}=3.3\). [While determining the value of Kp, pressure should be expressed in bar and the value of R must be in L bar K-1 mol)-1.]

Example 2. The concentration equilibrium constant, Kc for the reaction \(2 \mathrm{NO}_2 \rightleftharpoons \mathrm{N}_2 \mathrm{O}_4\) at 25° C is 216. Find the pressure equilibrium constant, Kp, of the reaction at the same temperature.
Solution:

Kp = Kc(RT)Δn

Δn =1- 2 =-1

∴ Kp =Kc(RT)-1

= \(\frac{K_c}{R T}\)

= \(\frac{216}{(0.08312) \times(298.15)}\).

Kp = 8.72. [R = 0.0312 L bar K-1 mol-1]

Example 3. The value of Kc for the production of phosgene, the reaction being \(\mathrm{CO}(\mathrm{g})+\mathrm{Cl}_2(\mathrm{~g}) \rightleftharpoons \mathrm{COCl}_2(\mathrm{~g})\), is 5 at 600 K. What are the equilibrium partial pressures ofthe gases if the initial pressures are as follows? \(p_{\mathrm{CO}}=0.265 \mathrm{bar}, p_{\mathrm{Cl}_2}=0.265 \mathrm{bar}, p_{\mathrm{COCl}_2}=0\).
Solution:

At first, Kp must be calculated from Kc.

Kp = Kc(RT)Δn

Here, Δn =1- 2 = -1

∴ Kp = 5 x (0.08312 x 600)-1 = 0.1

Let us now tabulate the partial pressures of all the gases. Suppose the partial pressure of COCl2 at equilibrium is x bar.

Partial pressures:

⇒ \(\begin{array}{lccc}p_{\mathrm{CO}} p_{\mathrm{CO}_2} p_{\mathrm{COCl}_2} \\
\text { Initial } 0.265 \text { bar } 0.265 \text { bar }  0\end{array}\)

⇒ \(\begin{array}{lccc}
\text { Change } -x -x -x \\
\text { Equilibrium } 0.265-x 0.265-x x
\end{array}\)

∴ \(\dot{K}_p=0.1=\frac{p_{\mathrm{COC}_2}}{p_{\mathrm{CO} p_{\mathrm{Cl}_2}}}=\frac{x}{(0.265-x)^2}\)

0.1 = \(\frac{x}{\left(0.07-0.53 x+x^2\right)}\)

Rearranging, we get

0.1 x2 – 0.053x + 0.007 = x

or 0.1 x2 – 0.053x – x + 0.007 = 0

or 0.1 x2 – 0.053 x + 0.007 = 0

or x = \(\frac{+1.053 \pm \sqrt{(1053)^2-4 \times 0.1 \times 0.007}}{2 \times 0.1}=\frac{1.053 \pm 1.0517}{0.2}=6.5 \times 10^{-3} \text { or } 10.52\)

x cannot be greater than 0.265 bar since even if all of CO and Cl2 react, \(\mathrm{p}_{\mathrm{COCl}_2}\), cannot be more than 0.265 bar.

Choosing x to be 6.5 x 10-3

at equilibrium, \(\mathrm{p}_{\mathrm{COCl}_2}\) =6.5 x 10-3 bar

PCO = 0-265 – 65 x 10-3 = 0.2585 bar = pa

Example 4. The standard Gibbs free energy at 298 K for the reactions are given.

  1. \(\mathrm{N}_2(\mathrm{~g})+\mathrm{O}_2(\mathrm{~g}) \rightleftharpoons 2 \mathrm{NO}(\mathrm{g}) \Delta \mathrm{G}^{\ominus}=173.2 \mathrm{~kJ}\)
  2. \(2 \mathrm{NO}(\mathrm{g})+\mathrm{O}_2(\mathrm{~g}) [latex] \rightleftharpoons 2 \mathrm{NO}_2(\mathrm{~g}) \Delta \mathrm{G}^{\ominus}=-69.7 \mathrm{~kJ}\)

Calculate Kp at 298 K for the following reaction \(\mathrm{N}_2(\mathrm{~g})+2 \mathrm{O}_2(\mathrm{~g}) \rightleftharpoons 2 \mathrm{NO}_2(\mathrm{~g})\)

Solution:

First, we must find \(\Delta G^{\ominus}\), for the reaction of interest.

⇒ \(\begin{aligned}
\mathrm{N}_2(\mathrm{~g})+\mathrm{O}_2(\mathrm{~g}) \rightleftharpoons 2 \mathrm{NO}(\mathrm{g}) \quad \Delta \mathrm{G}^{\ominus}=173.2 \mathrm{~kJ} \\
2 \mathrm{NO}(\mathrm{g})+\mathrm{O}_2(\mathrm{~g}) \rightleftharpoons 2 \mathrm{NO}_2(\mathrm{~g}) \quad \Delta G^{\ominus}=-69.7 \mathrm{~kJ} \\
\overline{\mathrm{N}_2(\mathrm{~g})+2 \mathrm{O}_2(\mathrm{~g}) \rightleftharpoons 2 \mathrm{NO}_2(\mathrm{~g}) \Delta G^{\ominus}=-173.2-69.7} \\
=0.103 .5 \mathrm{~kJ} \text {. } \\
\end{aligned}\)

Thus, \(\Delta G^{\ominus}\) for the reaction in question is the sum of \(\Delta G^{\ominus}\) for the other two reactions.

∴ \(\Delta G^{\ominus}\) = 103.5 kJ.

∴ \(K_p=e^{-\Delta G^{\ominus} / R T}\)

But \(\frac{-\Delta G^{\ominus}}{R T}=\frac{-103.5 \times 1000}{8.314 \times 298}=-4178\)

∴ \(\quad K_p=e^{-41.78}=7.16 \times 10^{-19}\).

Example 5. \(\Delta_f G^{\ominus}\) for ammonia gas is -16.5 kJ mol1 at 298 K. Find the equilibrium constant for the reaction \(\mathrm{N}_2(\mathrm{~g})+3 \mathrm{H}_2(\mathrm{~g}) \rightarrow 2 \mathrm{NH}_3(\mathrm{~g})\)
Solution:

2 mol of NH3 is produced in the reaction but the \(\Delta_f G^{\ominus}\) value given is for only 1 mol. \(\Delta_f G^{\ominus}\) for the formation of 2 mol of NH3 is equal to twice \(\Delta_f G^{\ominus}\)for 1 mol.

⇒ \(\Delta_f G^{\ominus}\) = 2 x -165 kJ = -33.0 kJ = -33,000 J.

Now, \(\frac{-\Delta_f G^{\ominus}}{R T}=\frac{33000}{8.314 \times 298}=13.32\)

∴ \(K_c=e^{-\Delta, G^4 / R T}=e^{+1332}=609,259.8=6.09 \times 10^5\).

Example 6. Kc for the hydrolysis of sucrose is 5.3 x 1012 at 298 K. Find \(\Delta G^{\ominus}\) for the process sucrose + H2O \(\rightleftharpoons\) glucose + fructose
Solution:

⇒ \(K_c=e^{-\Delta G^{\ominus} / R T}\)

ln \(K_c=-\Delta G^{\ominus} / R T\)

⇒ \(\Delta G^{\ominus}=-R T \ln K\)

= -8.314 x 298 x ln(53 x 1012) = -72,589.7 J.

∴ \(\Delta G^{\ominus}\)=-72.58 kJ.

Example 7. Consider the reaction \(\mathrm{PCl}_5(\mathrm{~g}) \rightleftharpoons \mathrm{PCl}_3(\mathrm{~g})+\mathrm{Cl}_2(\mathrm{~g})\). Kp =23.6 at 500K for the reaction. Calculate the equilibrium partial pressures of the reactants and products if the initial pressure of PCl5 is 0.56 bar and those of PCl3 and Cl2 are zero. How will the concentration of PCl5 and PCl3 change if more Cl2 is added once equilibrium is reached?
Solution:

Let us tabulate the partial pressure—initial, change and at equilibrium.

⇒ \(p_{\mathrm{PC}} p_{\mathrm{PC}_3} p_{\mathrm{C}_2}\)

⇒ \(\begin{array}{lccc}
\text { Initial } 0.56  0  0 \\
\text { Change } -x +x +x \\
\text { Equilibrium } 0.56-x  x  x
\end{array}\)

Putting these values in the expression for Kp,

⇒ \(K_p=\frac{p_{\mathrm{PCl}_3} p_{\mathrm{Cl}_2}}{p_{\mathrm{PCl}_5}}\)

23.6 = \(\frac{x \cdot x}{0.56-x}\)

Rearranging, we get 23.6(056- x)- x2 = 0

or x2– 23.6(056 -x) = 0

or x2 23.6 x 056 + 23.6x = 0

or x2+23.6x- 13.2 = 0.

∴ x = \(\frac{-23.6 \pm \sqrt{(23.6)^2+4 \times 13.2}}{2}\)

= \(\frac{-23.6 \pm 24.69}{2}=0.547 \text { or }-24.14\)

Taking the positive root, x = 0547, we get

⇒ \(p_{\mathrm{PCl}_5}\) = 0.56 -0547 = 0.013 bar

⇒ \(p_{\mathrm{PCl}_3}\) = 0.547 bar

⇒ \(p_{\mathrm{Cl}_2}\) = 0547 bar,

On adding Cl2 at equilibrium, the reaction will shift towards the left \(p_{\mathrm{PCl}_5}\) and will increase and \(p_{\mathrm{PCl}_3}\) will decrease.

Factors Affecting Equilibrium

In this section, we shall discuss what happens to a system at equilibrium when certain conditions are changed. We shall consider various changes, one by one.

  • What happens to the equilibrium when the temperature is changed, for example? Or what happens when the concentration of either the reactants or the products is changed?
  • In 1888, the French chemist Henri Louis Le Chatelier proposed a general law on the behaviour of a system in equilibrium.
  • This law, called Le Chatelier’s principle, states that if a system at equilibrium is subjected to a change which displaces if from the equilibrium, a net reaction will occur in a direction that counteracts the change.
  • This law is applicable to all physical and chemical equilibria. Let us now discuss in detail the various factors that can affect an equilibrium.

Change in concentration: Let us consider the reaction \(\mathrm{H}_2(\mathrm{~g})+\mathrm{I}_2(\mathrm{~g}) \rightleftharpoons 2 \mathrm{HI}(\mathrm{g})\) at equilibrium.

  • If more of the reactant H2 or I2, or both, are added to the reaction system, the state of equilibrium will be disturbed and the tendency of the system would be to counteract this change so that equilibrium is established again.
  • The only way this can happen is if more of the reactants react to form more of the products so that the ratio of the products of the concentrations the reaction products to that of the reactants, i.e., K, remains constant.
  • Another way of saying this is, if the concentration of the reactants is increased, the equilibrium shifts in the forward direction, or the forward reaction is favoured.
  • Applying Le Chatelier’s principle to a specific change (concentration), we may say, “When the concentration of any of the reactants or products in a reaction at equilibrium is changed, the composition of the equilibrium mixture changes so as to undo the effect of the concentration change.”

Let us study how equilibrium is re-established if we add H2 at equilibrium. The concentration of H2 increases and equilibrium is disturbed.

Basic Chemistry Class 11 Chapter 7 Equilibrium Effect Of Addition Of H2 bOn The Equilibrium The Reaction

A new equilibrium will be set up in which the concentration of H2 should be less than what it is after adding H2. But the initial concentration of H2 increases for the new equilibrium as shown in Figure.

  • As you can see from the figure, till the time tx, the system is at equilibrium with the concentrations of H2, I2 and HI being given by the intercept of the three curves on the y-axis.
  • At t1, when some H2 is added, its concentration increases as shown by the steep rise of the H2 curve. The equilibrium is disturbed and the system responds to the change by forming more HI, thus decreasing the concentration of H2 and I2 than at time t1. At t2 a new equilibrium is established.

Let us consider the reaction quotient for this reaction \(Q_c=\frac{[\mathrm{HI}]^2}{\left[\mathrm{H}_2\right]\left[\mathrm{I}_2\right]}\)

If hydrogen is added at equilibrium, the molar concentration of hydrogen increases and the equilibrium is disturbed. The denominator of the above equation increases and hence Qc decreases or becomes less than Kc.

  • Therefore, in order to make Qc equal to Kc the numerator must increase or the denominator must decrease. This means more HI should be formed in order to counteract the disturbance in equilibrium.
  • As a result more of H2 and I2 react to form more of HI to re-establish the equilibrium making the forward reaction more favourable.
  • What if the concentration of hydrogen iodide is increased? Again there will be a tendency for the reaction system to counteract this change.
  • This time, more hydrogen iodide will yield more hydrogen and iodine so that equilibrium is restored and Kc remains constant. That is to say, if the concentration of the products is increased, the equilibrium will shift in the reverse direction, or the reverse reaction will be favoured.

Changes in concentration (of either products or reactants) play an important role in the productivity of industrial processes. When ammonia is manufactured by the Haber process, for example, the equilibrium involved is the following.

⇒ \(3 \mathrm{H}_2(\mathrm{~g})+\mathrm{N}_2(\mathrm{~g}) \rightleftharpoons 2 \mathrm{NH}_3(\mathrm{~g})\)

The product, i.e., NH3, is continuously removed from the site of the reaction. This disturbs the equilibrium and the system counteracts the change by producing more ammonia.

  • Increasing the concentration of the reactants would have the same impact. But, obviously, the former is more economical.
  • The large-scale production of CaO from CaCO3 is another example: the constant removal of CO2 from the lime kiln drives the reaction to completion.
  • There are many applications of Le Chatelier’s principle in our everyday lives. For instance, on humid, cloudy days when there is hardly any breeze, we often dry clothes indoors, under the fan.
  • When we do this, we are using Le Chatelier’s principle without knowing it. The water present in the clothes and the water vapour in the air in the vicinity of the clothes reach an equilibrium, which we disturb by turning on the fan.
  • The artificial breeze (created by the fan) removes the water vapour from the immediate neighbourhood of the clothes and more water molecules from the clothes escape into the air to re-establish equilibrium.
  • For the same reason, one feels more comfortable under the fan on a sultry day.
  • A very vital life process depends on Le Chatelier’s principle. You know that haemoglobin (ITb) present in red blood corpuscles acts as a carrier of oxygen from the lungs to the tissues and as a carrier of carbon dioxide from the tissues to the lungs.

How does this happen? The blood that comes from the lungs and reaches the tissues has a high concentration of oxygen in comparison with the blood present in the tissues, where the partial pressure of oxygen is low.

This disturbs the equilibrium and in order to re-establish it, some of the oxyhaemoglobin (haemoglobin carrying oxygen) dissociates.

⇒ \(\mathrm{HbO}_2(\mathrm{~s}) \rightleftharpoons \mathrm{Hb}(\mathrm{s})+\mathrm{O}_2(\mathrm{~g})\)

When the blood returns from the tissues to the lungs, more oxyhaemoglobin is formed.

  • This happens because the concentration or partial pressure of oxygen in the lungs is high and to re-establish equilibrium, some of the haemoglobin in the blood returning from the tissues combines with oxygen to form oxyhaemoglobin.
  • The removal of carbon dioxide from the tissues by haemoglobin happens in a similar manner.
  • The partial pressure of carbon dioxide in the tissues is high, so some of it dissolves in the blood, which reaches the tissues from the lungs.

Carbon dioxide is released from the blood when it returns to the lungs, where the partial pressure of carbon dioxide is low.

⇒ \(\mathrm{CO}_2(\mathrm{~g})+\mathrm{H}_2 \mathrm{O}(\mathrm{l}) \rightleftharpoons \mathrm{H}_2 \mathrm{CO}_3(\mathrm{aq}) \rightleftharpoons \mathrm{H}^{+}(\mathrm{aq})+\mathrm{HCO}_3^{-}(\mathrm{aq})\)

The effect of concentration can also be demonstrated with the help of a simple experiment. You are familiar with the reaction

⇒ \(\underset{\text { yellow }}{\mathrm{Fe}^{3+}(\mathrm{aq})}+\underset{\text { colourless }}{\mathrm{SCN}^{-}(\mathrm{aq})} \rightleftharpoons \underset{\text { deep red }}{\mathrm{FeSCN}^{2+}(\mathrm{aq})}\)

⇒ \(K_c=\frac{\left[\mathrm{FeSCN}^{2+}(\mathrm{aq})\right]}{\left[\mathrm{Fe}^{3+}(\mathrm{aq})\right]\left[\mathrm{SCN}^{-}(\mathrm{aq})\right]}\)

If we take ferric nitrate solution and thiocyanate in a test tube in an appropriate proportion and shake, the colour changes from yellow to red due to the formation of FeSCN2+.

  • This happens gradually and the colour does not change once equilibrium is attained. We can shift the equilibrium by adding or removing the reactant or product.
  • If we add few more drops of thiocyanate solution in the test tube, Qc becomes less than Kc. Therefore, the forward reaction is favoured and more FeSCN2+ is formed until Qc = Kc and equilibrium is attained.
  • At first, the intensity of the red colour decreases and then increases to deep red, at equilibrium.
  • If on the other hand, we remove SCN by adding a reagent, say HgCl2(aq) which forms a stable complex—[Hg(SCN)4]2-—the equilibrium shifts in the reverse direction so that the concentration of FeSCN2+( decreases because it dissociates to replenish SCN.
  • A similar shift in the equilibrium will be observed if we add oxalic acid to the reaction mixture at equilibrium. Oxalic add reacts with Fe3+ to form a stable complex—[Fe(C2O4)3]3-.

Change in pressure: If one of the reactants or products in a reaction system is in the gaseous phase, pressure becomes an important factor governing shifts in equilibrium. In the case of a gas, increase in partial pressure amounts to increase in concentration.

  • Let us consider a heterogeneous system first. \(\mathrm{CO}_2(\mathrm{~g})+\mathrm{H}_2 \mathrm{O}(\mathrm{l}) \rightleftharpoons \mathrm{CO}_2(\mathrm{aq})\)
  • In this case increasing the partial pressure of CO2(g) would mean an increase in the number of moles of CO2 per unit volume (concentration).
  • According to Le Chatelier’s principle the system would try to counteract the change to re-establish equilibrium. The only way this is possible is if more CO2 dissolves in water. This is why the solubility’ of gases increases with increase in pressure.
  • Now consider a homogeneous system in which all the reactants and products are in the gaseous phase. The reaction could lead to an overall increase in the number of moles, if the number of moles of the products is more than that of the reactants.

It could lead to an overall decrease in the number of moles, if the number of moles of products is less than that of the reactants. There is a third possibility—there may be no net change in the number of moles. Let us consider these cases one by one.

1. The reaction involved in the manufacture of ammonia by the Haber process is as follows. \(3 \mathrm{H}_2(\mathrm{~g})+\mathrm{N}_2(\mathrm{~g}) \rightleftharpoons 2 \mathrm{NH}_3(\mathrm{~g})\)

  • This is obviously a case in which the forward reaction leads to a decrease in the number of moles.
  • In such a case, if the total pressure is increased, say by decreasing the volume, the number of moles per unit volume increases and the system tries to nullify the change by decreasing the number of moles per unit volume (we already know that pV = constant and p ∝ moles of the gas).
  • Since the forward reaction leads to a decrease in the number of moles, it is favoured and the equilibrium shifts in the forward direction.
  • In short, increasing the pressure leads to an increase in the yield of the products. Obviously, if pressure is decreased, the reverse reaction will be favoured.

This can also be understood by considering reaction quotient. Let[H2], [N2] and [NH3]be the molar concentrations at equilibrium in the reaction involved in the Haber process.

When the volume of the reaction mixture is halved the pressure as well as concentrations are doubled. Now, let us obtain the concentration quotient by replacing each value by its double.

⇒ \(Q_c=\frac{\left(2\left[\mathrm{NH}_3\right]\right)^2}{\left(2\left[\mathrm{H}_2\right]\right)^3\left(2\left[\mathrm{~N}_2\right]\right)}=\frac{1}{4} \frac{\left[\mathrm{NH}_3\right]}{\left[\mathrm{H}_2\right]\left[\mathrm{N}_2\right]}=\frac{K_c}{4}\)

As Qc < Kc, the forward reaction is favoured.

2. Let us consider an equilibrium in which the forward reaction involves an increase in the number of moles.

⇒ \(\mathrm{PCl}_5(\mathrm{~g}) \rightleftharpoons \mathrm{PCl}_3(\mathrm{~g})+\mathrm{Cl}_2(\mathrm{~g})\)

Here an increase in pressure will favour the backward reaction, while a decrease in pressure will favour the forward reaction.

3. In certain reactions, for example, the ones which follow, the number of moles of the products is equal to that of the reactants.

⇒ \(\mathrm{N}_2(\mathrm{~g})+\mathrm{O}_2(\mathrm{~g}) \rightleftharpoons 2 \mathrm{NO}(\mathrm{g})\)

⇒ \(\mathrm{H}_2(\mathrm{~g})+\mathrm{I}_2(\mathrm{~g})  \rightleftharpoons 2 \mathrm{HI}(\mathrm{g})\)

Changes in pressure have no impact on such reactions.

When we study the effect of change of pressure on a heterogenous system, the solid or liquid reactants or products are ignored because the volume (and concentration) of a solid or a liquid is nearly independent of pressure.

For instance, in the following reactions carbon and iodine are solids.

⇒ \(\mathrm{C}(\mathrm{s})+\mathrm{CO}_2(\mathrm{~g}) \rightleftharpoons 2 \mathrm{CO}(\mathrm{g})\)

⇒ \(\mathrm{H}_2(\mathrm{~g})+\mathrm{I}_2(\mathrm{~s} \rightleftharpoons 2 \mathrm{HI}(\mathrm{g})\)

When the pressure on the reaction mixture is increased, the reverse reaction is favoured because the number of moles of the gases decreases in the reverse direction.

Change in temperature: Temperature affects the equilibrium constant Kc. The nature and extent of the change in Kc or Kp of a reaction due to the change in temperature depends on the enthalpy change of the reaction.

Suppose the forward reaction in a chemical equilibrium is exothermic then the reverse reaction will, naturally, be endothermic. Production of ammonia by the Haber process involves an exothermic reaction.

⇒ \(\mathrm{N}_2(\mathrm{~g})+3 \mathrm{H}_2(\mathrm{~g}) \rightleftharpoons 2 \mathrm{NH}_3(\mathrm{~g}) ; \Delta_r H^{\ominus}=-9238 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

  • If such a reaction system attains equilibrium at a particular temperature and then heat is supplied to the system so that the temperature increases, the equilibrium gets disturbed.
  • According to Le Chatelier’s principle, the system then tries to counteract the change in temperature and the only way it can do this is by absorbing heat.
  • In other words, the equilibrium shifts to the left and the endothermic or reverse reaction is favoured.
  • If, on the other hand, the temperature is decreased, the equilibrium shifts forward, favouring the forward or exothermic reaction, thus leading to high yield of ammonia.

In this context, refer to the experiment in Figure. NO2 dimerises into N2O4 according to the following reaction.

⇒ \(2 \mathrm{NO}_2(\mathrm{~g}) \rightleftharpoons \mathrm{N}_2 \mathrm{O}_4(\mathrm{~g}) ; \Delta_r H^{\ominus}=-57.2 \mathrm{~kJ} \mathrm{~mol}^{-1} \text {. }\)

Here the formation of N2O4 is favoured at a low temperature (ice bath) whereas that of NO2 is favoured at high temperature (hot water).

This reaction is exothermic. On increasing the temperature, in order to relieve the stress of added heat, the system will tend to absorb this heat and the reverse reaction will be favoured.

  • Since NO2 is brown and N2O4 is colourless, the effect of increasing temperature on this reaction may easily be seen.
  • Decreasing the temperature, say by immersing the vessel containing a mixture of NO2 and N2O4 in an ice bath, favours the forward reaction.
  • When the dissolution of a solid in water (or any other solvent) is accompanied by the absorption (endothermic) or release (exothermic) of heat, its solubility changes with temperature.
  • When NH4Cl dissolves in water, for example, heat is absorbed. The solubility of NH4Cl increases with temperature because an increase in temperature favours the endothermic process.
  • CaCl2, on the other hand, dissolves in water with the evolution of heat. The solubility of CaCl2, thus, decreases with increase in temperature.

The solubility of NaCl is almost unaffected by changes in temperature because there is very little heat change during the dissolution of NaCl.

Effect of a catalyst: A catalyst merely changes the time taken to reach the state of equilibrium, it does not change the equilibrium constant. It lowers the activation energy of the forward and reverse reactions by the same amount.

  • The activation energy is the energy barrier that has to be overcome for the reaction to proceed. If a reaction mixture is not at equilibrium, a catalyst accelerates the rate at which equilibrium is reached but it does not affect the composition of the equilibrium mixture.
  • Hence it does not appear in the balanced chemical equation or in the equilibrium constant expression.
  • It increases the rates of the forward and reverse reactions to the same extent. Nonetheless, the fact that a catalyst decreases the time taken to reach equilibrium is very important for many industrial processes.
  • For instance, iron-molybdenum is used as a catalyst in the production of ammonia from nitrogen and hydrogen because at low temperature, the rate of the reaction is very slow and at high temperature, the reverse reaction is favoured.
  • The optimum conditions of temperature and pressure for the synthesis of NH3 using a catalyst are 400-500°C and 130-300 atm.

In the manufacture of sulphuric acid by the contact process, the conversion of SO2 to SO3 is very important.

⇒ \(2 \mathrm{SO}_2(\mathrm{~g})+\mathrm{O}_2(\mathrm{~g}) \rightleftharpoons 2 \mathrm{SO}_3(\mathrm{~g}) ; K_{\mathrm{c}}=1.7 \times 10^{26}\)

The magnitude of the equilibrium constant indicates that the reaction should reach completion.

However, the oxidation of SO2 is very slow but in the presence of the catalyst V2O5 (vanadium pentoxide), the rate of the reaction increases. But if a reaction has a very low Kc, a catalyst would be of no help.

Effect of inert gas: While studying the effect of pressure, we have considered changes that result from a change in volume (as volume and pressure are inversely proportional).

  • But what if we keep the volume constant and increase the pressure by adding an inert gas. The inert gas is added just to change the pressure—it does not take part in the reaction.
  • We could also think of adding an inert gas at constant pressure. If an inert gas is added to a reaction system containing reactants and products in the gaseous phase at equilibrium, the effect it will have on the equilibrium will depend on whether it is added at constant volume or at constant pressure.
  • If the equilibrium is readied at constant volume, i.e., in a closed vessel, the addition of the inert gas will not change the molar concentrations (or partial pressures) of the reactants and products. Consequently, the state of equilibrium will not be affected.

If an inert gas is added to a system at equilibrium at constant pressure, the volume will increase.

  • An increase in volume will lead to a decrease in the molar concentrations (or partial pressures) of all the reactants and products.
  • If the reaction is such that the number of moles of the reactants is the same as that of the products, a decrease in the molar concentrations of all the reactants and products will make no difference to the equilibrium constant since it is the ratio of the product of concentrations of the products to that of the reactants.
  • (We came across a similar situation while studying the effect of a change in pressure on an equilibrium.) If, however, the number of moles of the reactants is not the same as that of the products, an increase in the total volume will disturb the equilibrium.

Suppose the number of moles of tire reactants is less than that of the products, as in the following equilibrium.

⇒ \(\mathrm{PCl}_5(\mathrm{~g}) \rightleftharpoons \mathrm{PCl}_3(\mathrm{~g})+\mathrm{Cl}_2(\mathrm{~g})\)

In such a case, a decrease in the molar concentrations (due to increase in total volume) of all the products and reactants will lead the system to counteract the change and the equilibrium will shift in the direction of increasing the number of moles, i.e., the forward direction.

  • We can look at it in another way. At equilibrium \(K_{\mathrm{c}}=\frac{\left[\mathrm{PCl}_3\right]\left[\mathrm{Cl}_2\right]}{\left[\mathrm{PCl}_5\right]}\)
  • If the concentrations of all the reactants and products decrease, the numerator will decrease more since there are two concentration terms in the numerator and Kc will decrease.
  • For Kc to remain constant, the equilibrium will shift in the direction of an increase in the concentrations of products and decrease in the concentration of the reactants, i.e., in the forward direction.

Similarly, if a reaction is such that the number of moles of the reactants is more than that of the products, an increase in volume will push the equilibrium in the reverse direction.

⇒ \(2 \mathrm{SO}_2(\mathrm{~g})+\mathrm{O}_2(\mathrm{~g}) \rightleftharpoons 2 \mathrm{SO}_3(\mathrm{~g})\)

For example, for the reaction an increase in volume will favour the reverse reaction. In general, the addition of an inert gas at constant pressure (to a system at equilibrium) makes the equilibrium shift in the direction that increases the number of moles.

Ionic Equilibrium

When certain compounds like NaCl, KCl or NaOH are dissolved in water, the resultant solution is a good conductor of electricity. The solution of NaCl contains Na+ and Cl ions surrounded by water.

⇒ \(\mathrm{NaCl}(\mathrm{s}) \stackrel{\mathrm{H}_2 \mathrm{O}}{\longrightarrow} \mathrm{Na}^{+}(\mathrm{aq})+\mathrm{Cl}^{-}(\mathrm{aq})\)

⇒ \(\mathrm{KCl}(\mathrm{s}) \stackrel{\mathrm{H}_2 \mathrm{O}}{\longrightarrow} \mathrm{K}^{+}(\mathrm{aq})+\mathrm{Cl}^{-}(\mathrm{aq})\)

  • The electrostatic attraction between the ion and water (ion-dipole interaction) results in the dissolution of such ionic compounds in water. The figure shows various interactions in the solution of KCI in water.
  • When KCl is dissolved in water, the positive end of the polar water molecule (hydrogen) is attracted by negative chloride ions at the surface of the solid, similarly, the negative end of the water molecule (oxygen) is attracted by the positive potassium (K+) ions.
  • The water molecules penetrate between K+ and Cl and surround them, thus reducing the strong interionic forces holding them together and allowing them to move as hydrated ions.
  • This process is called ionisation, i.e., the process of breaking up of a compound into its ions in a solution.
  • Though the exact definition of ionisation differs from that of dissociation, we have used both the terms to refer to the breaking up of an ionic substance into its constituent into ions for the sake of convenience.
  • Michael Faraday in 1824 classified substances into electrolytes and nonelectrolytes. Nonelectrolytes do not conduct electricity.
  • Electrical conductivity requires the existence of charged particles and in a solution of nonelectrolyte, the solute molecules retain their unbroken identity.

Basic Chemistry Class 11 Chapter 7 Equilibrium Dissolution Of Potassium Chloride In Water

That is why though sucrose dissolves in water, it is not a conductor of electricity.

⇒ \(\mathrm{C}_{12} \mathrm{H}_{22} \mathrm{O}_{11}(\mathrm{~s}) \stackrel{\mathrm{H}_2 \mathrm{O}}{\longrightarrow} \mathrm{C}_{12} \mathrm{H}_{22} \mathrm{O}_{11}(\mathrm{aq})\)

Not all electrolytes conduct electricity with equal ease. Faraday categorised electrolytes as strong and weak. For instance, sodium chloride is a strong electrolyte whereas acetic acid is weak electrolyte.

Ionisation: All electrolytes do not ionise to the same extent. The extent to which an electrolyte ionises is called its degree of dissociation.

  • More accurately, the fraction of the total number of molecules of the electrolyte in the solution which dissociates into ions is called the degree of dissociation of the electrolyte.
  • A truly ionic compound dissociates completely in solution. For example, if 1 mol of NaCl is dissolved in 1 L of water, the solution will contain 1 mol of Na+ and 1 mol of Cl ions.
  • Compounds which dissociate completely in solution are called strong electrolytes.
  • Polar covalent compounds may also produce ionic solutions or solutions which conduct electricity.
  • Depending on the degree of ionisation or dissociation, such solutes are strong electrolytes or weak electrolytes.

H2SO4, for example, dissociates completely in solution and is a strong electrolyte. Equations for the ionisation of strong electrolytes are written with a single arrow.

⇒ \(\mathrm{HCl}(\mathrm{aq}) \longrightarrow \mathrm{H}^{+}(\mathrm{aq})+\mathrm{Cl}^{-}(\mathrm{aq})\)

⇒ \(\mathrm{NaOH}(\mathrm{aq}) \longrightarrow \mathrm{Na}^{+}(\mathrm{aq})+\mathrm{OH}^{-}(\mathrm{aq})\)

Weak electrolytes, like NH3 and CH3COOH (acetic acid), on the other hand, do not dissociate completely. Their solutions arc weakly or partly ionised.

In such cases there is an equilibrium between the un-ionised molecules of the electrolyte and the ions in solution. Consequently, the ionisation of such an electrolyte is represented with a double arrow.

⇒ \(\mathrm{NH}_3(\mathrm{aq})+\mathrm{H}_2 \mathrm{O}(\mathrm{l}) \rightleftharpoons \mathrm{NH}_4^{+}(\mathrm{aq})+\mathrm{OH}^{-}(\mathrm{aq})\)

⇒ \(\mathrm{CH}_3 \mathrm{COOH}(\mathrm{aq})+\mathrm{H}_2 \mathrm{O}(\mathrm{l}) \rightleftharpoons \mathrm{H}_3 \mathrm{O}^{+}(\mathrm{aq})+\mathrm{CH}_3 \mathrm{COO}^{-}(\mathrm{aq})\)

  • The degree of ionisation of weak electrolytes depends upon the nature of the electrolyte, temperature and dilution.
  • We will not concern ourselves with the ionisation of strong electrolytes because such reactions proceed to completion and in this chapter we are examining reactions from the point of view of equilibrium.

Ionisation of weak electrolytes: An equilibrium between the (un-ionised) molecules of a weak electrolyte and its ions in solution is called an ionic equilibrium.

  • If C is the concentration of an electrolyte in a solution, and a is the degree of dissociation, the number of moles of the electrolyte that dissociate is Cα.
  • The degree of dissociation (α) of an electrolyte is the fraction of the total number of molecules of the electrolyte (in the solution) that ionise at equilibrium.
  • For example, if the value of α = 0.5 or 1/2, it means that out of 1 mole of the electrolyte, 0.5 mol has ionised or there is 50% ionisation.
  • So, if C is the concentration of the electrolyte, what is the concentration that is ionised? It is Cα.

Let us say that the concentration of some electrolyte XY in an aqueous solution is C and its degree of dissociation is a. At equilibrium, the ionisation of the electrolyte can be represented by

⇒ \(\mathrm{XY}(\mathrm{aq}) \rightleftharpoons \mathrm{X}^{+}(\mathrm{aq})+\mathrm{Y}^{-}(\mathrm{aq})\)

At equilibrium, the concentrations of X+ and Y must each be Cα. And the concentration of XY must be C -Cα = C(1 – α). (The concentration of X+ and Yare equal to the concentration of XY dissociated.)

Applying the law of chemical equilibrium

K = \(\frac{\left[\mathrm{X}^{+}(\mathrm{aq})\right]\left[\mathrm{Y}^{-}(\mathrm{aq})\right]}{[\mathrm{XY}(\mathrm{aq})]}\)

= \(\frac{(\mathrm{C} \alpha)(\mathrm{C} \alpha)}{C(1-\alpha)}=\frac{C \alpha^2}{1-\alpha}\).

For a weak electrolyte, a is so small compared to 1 that it can be neglected in approximate calculations. Then K = Cα2

or \(\quad \alpha^2=\frac{K}{C}\)

or \(\alpha=\sqrt{\frac{K}{C}}\)

⇒ \(\quad \alpha \propto \sqrt{\frac{1}{C}}\) (because K is a constant)

  • Therefore, for a weak electrolyte, the degree of ionisation is inversely proportional to the square root of the molar concentration.
  • This is known as Ostwald’s dilution law. As C approaches zero, or dilution approaches infinity, the degree of dissociation approaches unity.
  • This means that the ionisation of a weak electrolyte will be more in a dilute solution than in a concentrated solution.

Example: Calculate the degree of dissociation of a 0.02-M solution of formic acid, given that K = 1.9 x 10-4.
Solution:

The dissociation of formic acid in water can be represented as follows.

⇒ \(\mathrm{HCOOH}+\mathrm{H}_2 \mathrm{O} \rightleftharpoons \mathrm{H}_3 \mathrm{O}^{+}+\mathrm{HCOO}^{-}\)

Let α be the degree of dissociation. Then the concentrations of the ionic species at equilibrium will be as follows:

[H3O]+ = 0.02α

[HCOO] = 0.02α

[HCOOH] = 0.02 (1 – α) = 0.02, assuming a to be negligible compared to 1.

Now K = \(\frac{(0.02 \alpha)(0.02) \alpha}{0.02}=1.9 \times 10^{-4}\) (given)

or \(\alpha^2=\frac{1.9 \times 10^{-4}}{0.02}\)

or \(\alpha=\sqrt{95} \times 10^{-2}=9.75 \times 10^{-2}\).

The degree of dissociation of formic acid = 0.0097.

Acid Base Equilibrium

Acids and bases are also electrolytes. You are quite familiar with some properties of acids and bases. You know, for example, that adds turn blue litmus red and that bases (also called alkalis) turn red litmus blue.

You know that adds react with active metals like zinc and magnesium to liberate hydrogen. You know that acids are sour and corrosive and that bases are bitter and slippery. And you also know that adds and bases neutralise each other.

This is essentially the concept chemists had of acids and bases in the olden days. You could call this concept, based on certain properties of adds and bases, the dassical concept.

Arrhenius theory: In 1884, S A Arrhenius, a Swedish chemist, proposed a new definition of acids and bases. It was based on his general theory of ionisation of electrolytes, which basically said that an electrolyte, on dissolution in water, dissociated into positively and negatively charged ions.

Arrhenius defined an acid as a substance which produces hydrogen ions (H+) when mixed with water. Bases, according to Arrhenius, are substances which produce hydroxyl ions (OH) when mixed with water. And the neutralisation of adds and bases is a reaction between H+ and OH ions.

⇒ \(\mathrm{H}^{+}(\mathrm{aq})+\mathrm{OH}^{-}(\mathrm{aq}) \rightleftharpoons \mathrm{H}_2 \mathrm{O}(\mathrm{l})\)

Strong acids like HCl, HNO3 and H2SO4 dissociate completely when dissolved in water.

⇒ \(\mathrm{HCl} \stackrel{\text { water }}{\longrightarrow} \mathrm{H}^{+}+\mathrm{Cl}^{-}\)

⇒ \(\mathrm{HNO}_3 \stackrel{\text { water }}{\longrightarrow} \mathrm{H}^{+}+\mathrm{NO}_3\)

Weak acids like acetic acid, carbonic acid and phosphoric acid dissociate only to a small extent when dissolved in water.

⇒ \(\mathrm{CH}_3 \mathrm{COOH} \stackrel{\text { water }}{\rightleftharpoons} \mathrm{CH}_3 \mathrm{COO}^{-}+\mathrm{H}^{+}\)

⇒ \(\mathrm{H}_2 \mathrm{CO}_3 \stackrel{\text { water }}{\rightleftharpoons} 2 \mathrm{H}^{+}+\mathrm{CO}_3^{2-} \)

⇒ \(\mathrm{H}_3 \mathrm{PO}_4 \stackrel{\text { water }}{\rightleftharpoons}{=} 3 \mathrm{H}^{+}+\mathrm{PO}_4^{3-}\)

Bases like NaOH and KOH are strong bases because they dissociate completely.

⇒ \(\mathrm{NaOH} \stackrel{\text { water }}{\longrightarrow} \mathrm{Na}^{+}+\mathrm{OH}^{-}\)

⇒ \(\mathrm{KOH} \stackrel{\text { water }}{\longrightarrow} \mathrm{K}^{+}+\mathrm{OH}^{-}\)

Substances like NH4OH, Ca(OH)2 and Al(OH)3 dissociate to a small extent and are called weak bases.

⇒ \(\mathrm{NH}_4 \mathrm{OH} \stackrel{\text { water }}{\rightleftharpoons} \mathrm{NH}_4^{+}+\mathrm{OH}^{-}\)

⇒ \(\mathrm{Ca}(\mathrm{OH})_2 \stackrel{\text { water }}{\rightleftharpoons} \mathrm{Ca}^{2+}+2 \mathrm{OH}^{-}\)

⇒ \(\mathrm{Al}(\mathrm{OH})_3 \stackrel{\text { water }}{\rightleftharpoons} \mathrm{Al}^{3+}+3 \mathrm{OH}^{-}\)

  • According to Arrhenius, acids produce H+ ions and bases produce OH ions when dissolved in water.
  • Actually, the H+ ion, formed by the loss of an electron from the hydrogen atom, is only an unshielded proton.
  • This proton cannot exist independently since its charge density is very high.

In an aqueous solution, it binds itself with one of the two lone pairs available on the water molecule to form a hydronium ion (H3O+).

⇒ \(\mathrm{H}^{+}+\mathrm{H}_2 \mathrm{O} \longrightarrow \mathrm{H}_3 \mathrm{O}^{+}\)

For the sake of convenience, the hydrated proton (hydronium ion) is represented as H+ (aq). In an aqueous solution, the H+ and OH ions exist as hydrated ions.

Limitations of Arrhenius’s theory This concept of acids and bases could explain neutralisation, salt hydrolysis and the strength of adds and bases, but it had a few drawbacks.

  1. This theory cannot explain the basic nature of substances like NH3, Na2CO3 and CaO, which do not possess a hydroxyl (OH) group. Nor cannot it explain the basic nature of CO2, SO2, SO3, etc., which do not contain hydrogen.
  2. Another limitation of this theory is its inability to explain reactions between acidic and basic substances in the absence of water. Two examples follow.

⇒ \(\mathrm{SO}_3(\mathrm{~g})+\mathrm{CaO}(\mathrm{s}) \longrightarrow \mathrm{CaSO}_4(\mathrm{~s})\)

⇒ \(\mathrm{NH}_3(\mathrm{~g})+\mathrm{HCl}(\mathrm{g}) \longrightarrow \mathrm{NH}_4 \mathrm{Cl}(\mathrm{s})\)

Bronsted-Lowry theory: In 1923, a Danish chemist, I N Bronsted, and a British chemist, T M Lowry, independently put forward a more general theory of acids and bases.

According to this theory, an acid is a proton donor and a base is a proton acceptor. This definition has some very interesting implications. To understand these, let us consider some examples.

⇒ \(\mathrm{HCl}+\mathrm{H}_2 \mathrm{O} \rightleftharpoons \mathrm{H}_3 \mathrm{O}^{+}+\mathrm{Cl}^{-}\)

⇒ \(\mathrm{CH}_3 \mathrm{COOH}+\mathrm{H}_2 \mathrm{O}  \rightleftharpoons \mathrm{H}_3 \mathrm{O}^{+}+\mathrm{CH}_3 \mathrm{COO}^{-}\)

⇒ \(\mathrm{CO}_3^{2-}+\mathrm{H}_2 \mathrm{O} \rightleftharpoons \mathrm{HCO}_3^{-}+\mathrm{OH}^{-}\)

⇒ \(\mathrm{NH}_3+\mathrm{H}_2 \mathrm{O} \rightleftharpoons \mathrm{NH}_4^{+}+\mathrm{OH}^{-}\)

⇒ \(\mathrm{NH}_4^{+}+\mathrm{H}_2 \mathrm{O}  \rightleftharpoons \mathrm{H}_3 \mathrm{O}^{+}+\mathrm{NH}_3\)

⇒ \(\mathrm{HCl}+\mathrm{NH}_3 \rightleftharpoons \mathrm{NH}_4^{+}+\mathrm{Cl}^{-}\)

Consider the first two reactions. HCl is an acid because it donates one proton to become Cl and H2O is a base because it accepts a proton to become H3O+.

  • Similarly, in the second equation, CH3COOH is an acid and H2O is a base. Now consider the next two reactions. In these, water acts as an acid.
  • It donates a proton to become OH. This means that the same substance can act as an acid or a base. Such substances which can act both as an acid and a base are called amphoteric.
  • This definition of acids and bases does not restrict itself to neutral molecules. Even ions can act as acids or bases. For example, CO32- in the third equation acts as a base by accepting a proton.

Nor is it necessary for an acid-base reaction to take place in an aqueous medium. For example, in the sixth equation HCI acts as an acid by donating a proton and NH3 acts as a base by accepting a proton.

  • The presence of the hydroxyl group (OH) is not necessary for a substance to act as a base. Similarly, a substance can act as an acid even if it does not contain hydrogen.
  • For example, NH3 acts as a base in the fourth reaction. All that is necessary for an acid-base reaction is that one substance should be able to donate a proton and the other should be able to accept the proton.
  • A corollary is that a substance which acts as an acid in the presence of a proton-acceptor, will not act in the same way in the absence of a proton-acceptor.
  • For example, though acetic acid acts as an acid in an aqueous solution (second equation), it does not do so in a benzene solution because benzene is not a proton-acceptor.

The Bronsted-Lowry concept of acids and bases has a very interesting consequence. When a substance acts as an acid it donates a proton. But on donating a proton, it becomes a base because it becomes capable of accepting a proton. Let us take the case of NH3 and H2O.

⇒ \(\mathrm{H}_2 \mathrm{O}+\mathrm{NH}_3 \rightleftharpoons \mathrm{NH}_4^{+}+\mathrm{OH}^{-}\)

Here water acts as an acid by donating a proton and becomes OH. Now OH can accept a proton, so it can act as a base.

  • Similarly, NH3 acts as a base by accepting a proton and becomes NH+4, and NH+4 has a proton to donate, so it is an acid.
  • Now do you see how every acid turns into a base by donating a proton, and every base turns into an add by accepting a proton.
  • The base formed when an acid donates a proton is called the conjugate base of the add and the acid formed when a base accepts a proton is called the conjugate acid of the base.

In the reaction between NH3 and H2O, OH is the conjugate base of H2O and NH4 is the conjugate acid of NH3.

Thus, there are two acid-base pairs in the reaction and these are called conjugate acid-base pairs. A few examples will make this clearer.

⇒ \(\text { Acid } 1+\text { base } 2 \rightleftharpoons \text { acid } 2+\text { base } 1\)

⇒ \(\mathrm{HCl}+\mathrm{H}_2 \mathrm{O} \rightleftharpoons \mathrm{H}_3 \mathrm{O}^{+}+\mathrm{Cl}^{-}\)

⇒ \(\mathrm{H}_2 \mathrm{O}+\mathrm{CO}_3^{2-} \rightleftharpoons \mathrm{HCO}_3^{-}+\mathrm{OH}^{-}\)

⇒ \(\mathrm{HCO}_3^{-}+\mathrm{NH}_3 \rightleftharpoons \mathrm{NH}_4^{+}+\mathrm{CO}_3^{2-}\)

⇒ \(\mathrm{NH}_4^{+}+\mathrm{CH}_3 \mathrm{COO}^{-} \rightleftharpoons \mathrm{CH}_3 \mathrm{COOH}+\mathrm{NH}_3\)

  • A strong add has a strong tendency to donate a proton but its conjugate base does not have an equally strong tendency to accept a proton.
  • The conjugate base of a strong acid is weak and the conjugate base of a weak acid is strong. Also, if two acids are mixed, the weaker acid acts as a base.

⇒ \(\underset{\text { (perchloric acid) }}{\mathrm{HClO}_4}+\mathrm{H}_2 \mathrm{SO}_4 \rightleftharpoons \mathrm{H}_3 \mathrm{SO}_4^{+}+\mathrm{ClO}_4^{-}\)

In this reaction, H2SO4 acts as a base with respect to the stronger acid, HClO4. Thus, H2SO4 is amphoteric or amphiprotic. Certain acids can lose two or three protons. They are called diprotic (H2SO4) or triprotic (H3PO4) adds.

Strengths of acids and bases: The strength of an acid or base depends upon its readiness to lose or gain a proton. Experimentally, the strength of an acid or base is measured by its ionisation constant or dissociation constant.

  • A strong acid like HCl, HNO3, H2So4, HBr, HI and HClO4 dissociates almost completely in water. This means the molar concentration of H3O+ ions in an aqueous solution of a strong acid is almost the same as that of the acid itself.
  • But this is not the case for weak acids and we can compare the strengths of such acids if we know the dissociation constants at a particular temperature.
  • The equilibrium constant for the dissociation of an acid is called its acid dissociation constant or acid ionisation constant and is denoted by Ka.
  • Similarly, the base dissociation or base ionisation constant is Kb.

Consider a weak acid HA and suppose its dissociation equilibrium can be represented as follows.

⇒ \(\mathrm{HA}(\mathrm{aq})+\mathrm{H}_2 \mathrm{O}(\mathrm{l}) \rightleftharpoons \mathrm{H}_3 \mathrm{O}^{+}(\mathrm{aq})+\mathrm{A}^{-}(\mathrm{aq})\)

At equilibrium, \(K_{\mathrm{a}}=\frac{\left[\mathrm{H}_3 \mathrm{O}^{+}\right]\left[\mathrm{A}^{-}\right]}{[\mathrm{HA}]\left[\mathrm{H}_2 \mathrm{O}\right]}\). Here [H2O] = 1 as it is in a very large quantity and remains almost unchanged.

Suppose the initial concentration of HA is C mol L-1 and its degree of dissociation is α. Then

⇒ \(K_{\mathrm{a}}=\frac{(C \alpha)^2}{C(1-\alpha)}=C \alpha^2\) (neglecting a in comparison with 1)

or \(\alpha =\sqrt{\frac{K_{\mathrm{a}}}{C}}\).

Thus, the degree of dissociation is inversely proportional to the square root of the concentration. Suppose we consider two acids of equimolar concentration (i.e., C is the same for both) and degrees of dissociation α1 and α2.

Then, \(\frac{\alpha_1}{\alpha_2}=\sqrt{\frac{K_{a_1}}{K_{a_2}}}\)

However, the degree of dissociation of an acid is a measure of its strength.

Therefore, \(\frac{\text { strength of acid }}{\text { strength of } \text { acid }_2}=\sqrt{\frac{K_{\mathrm{a}_1}}{K_{\mathrm{a}_2}}} \text {. }\)

  • So, we may say that at a given temperature, the ionisation or dissociation constant of an add is the measure of
    its strength.
  • As you already know, the equilibrium constant is a dimensionless quantity so Ka is also without any unit.
  • Concentrations of all the speeds at ionisation equilibrium are measured with respect to the standard state concentration of 1 M.
  • The larger the value of Ka, the stronger is the acid. Weak acids, however, show a great variation in their strengths. For instance, both formic add and hydrocyanic add are weak but their ionisation constants are very different.

Basic Chemistry Class 11 Chapter 7 Equilibrium Ionisation Constants For Some Weak Acids At 298K

Example: The ionisation constants of formic acid (HCOOH) and hydrocyanic acid (HCN) are 1.77 x 1024 and  4.9 x 10-10 respectively. Compare their strengths.
Solution:

\(\frac{\text { Strength of formic acid }}{\text { Strength of hydrocyanic acid }}=\sqrt{\frac{K_{\mathrm{a}}(\mathrm{HCOOH})}{K_{\mathrm{a}}(\mathrm{HCN})}}=\sqrt{\frac{1.77 \times 10^{-4}}{4.9 \times 10^{-10}}}=601 \text {. }\)

Therefore, formic acid is 601 times stronger than hydrocyanic acid.

The relative strengths of bases can be determined in a similar manner. First, let us consider the equilibria involved in solutions of weak bases. If B is the base, the equilibrium may be represented as

⇒ \(\mathrm{B}(\mathrm{aq})+\mathrm{H}_2 \mathrm{O} \rightleftharpoons \mathrm{BH}^{+}(\mathrm{aq})+\mathrm{OH}^{-}(\mathrm{aq})\)

The equilibrium constant is Kb and is called base dissociation constant. The higher the value of Kb, the stronger is the base.

⇒ \(K_{\mathrm{b}}=\frac{\left[\mathrm{BH}^{+}\right]\left[\mathrm{OH}^{-}\right]}{[\mathrm{B}]}\)

If C is the molar concentration of a base and α is its degree of dissociation \(\alpha=\sqrt{\frac{K_b}{C}},\)

where Kb is the dissociation constant of the base.

If we take two bases b1 and b2 of equimolar concentrations and dissociation constants \(K_{b_1} \text { and } K_{b_2}\) then

∴ \(\frac{\text { strength of base }}{\text { strength of base }{ }_2}=\sqrt{\frac{K_{b_1}}{K_{b_2}}} \text {. }\)

Basic Chemistry Class 11 Chapter 7 Equilibrium Ionisation Constats For Some Weak Bases At 298K

Relation between Ka and Kb We have seen that the strength of an acid (or a base) is expressed by its Ka (or Kb). For a conjugate acid-base pair, the two equilibrium constants are related in a simple manner.

Let us consider the case of the acid-base pair NH4 and NH3. The equilibrium for both acid and base may be represented as

⇒ \(\mathrm{NH}_4^{+}(\mathrm{aq})+\mathrm{H}_2 \mathrm{O}(\mathrm{l}) \rightleftharpoons \mathrm{H}_3 \mathrm{O}^{+}(\mathrm{aq})+\mathrm{NH}_3(\mathrm{aq})\)

∴ \(K_{\mathrm{a}}=\frac{\left[\mathrm{H}_3 \mathrm{O}^{+}\right]\left[\mathrm{NH}_3\right]}{\left[\mathrm{NH}_4^{+}\right]}=5.6 \times 10^{-10}\)

∴ \(\mathrm{NH}_3(\mathrm{aq})+\mathrm{H}_2 \mathrm{O}(\mathrm{l}) \rightleftharpoons \mathrm{NH}_4^{+}(\mathrm{aq})+\mathrm{OH}^{-}(\mathrm{aq})\)

∴ \(K_{\mathrm{b}}=\frac{\left[\mathrm{NH}_4^{+}\right]\left[\mathrm{OH}^{-}\right]}{\left[\mathrm{NH}_3\right]}=1.8 \times 10^{-5}\).

On adding the two, we get the net reaction as \(2 \mathrm{H}_2 \mathrm{O}(\mathrm{l}) \rightleftharpoons \mathrm{H}_3 \mathrm{O}^{+}(\mathrm{aq})+\mathrm{OH}^{-}(\mathrm{aq})\)

The sum of the two reactions is simply the dissociation of water, and its equilibrium constant,

⇒ \(K_{\mathrm{w}}=\frac{\left[\mathrm{H}_3 \mathrm{O}^{+}\right]\left[\mathrm{OH}^{-}\right]}{\left[\mathrm{H}_2 \mathrm{O}\right]^2}\)

or \(K_{\mathrm{w}}=\left[\mathrm{H}_3 \mathrm{O}^{+}\right]\left[\mathrm{OH}^{-}\right]=1.0 \times 10^{-14}\).

The equilibrium constant for the net reaction is equal to the product of the equilibrium constants for the individual reactions.

⇒ \(K_{\mathrm{a}} \times K_{\mathrm{b}}=\frac{\left[\mathrm{H}_3 \mathrm{O}^{+}\right]\left[\mathrm{NH}_3\right]}{\left[\mathrm{NH}_4^{+}\right]} \times \frac{\left[\mathrm{NH}_4^{+}\right]\left[\mathrm{OH}^{-}\right]}{\left[\mathrm{NH}_3\right]}=\left[\mathrm{H}_3 \mathrm{O}^{+}\right]\left[\mathrm{OH}^{-}\right]=K_w,\)

i. e., (5.6 x 10-10) x (18 x 10-5) = 10 x 10-14. The ionic product of water (Kw) is discussed later in the chapter.

Whenever an equation can be written as a sum of two or more equations, the equilibrium constant for the net reaction is the product of the equilibrium constants of all the individual reactions.

Di- and polybasic acids and di- and polyacidic bases Polybasic or polyprotic acids are those which have more than one ionisable proton. For example, sulphuric acid has two and phosphoric acid has three ionisable protons. The ionisation reactions for a dibasic acid like H2SO4 can written as

⇒ \(\mathrm{H}_2 \mathrm{SO}_4(\mathrm{aq})+\mathrm{H}_2 \mathrm{O}(\mathrm{l})\rightleftharpoons \mathrm{H}_3 \mathrm{O}^{+}(\mathrm{aq})+\mathrm{HSO}_4^{-}(\mathrm{aq}) \)

⇒ \(\mathrm{HSO}_4^{-}(\mathrm{aq}) \rightleftharpoons \mathrm{H}_3 \mathrm{O}^{+}(\mathrm{aq})+\mathrm{SO}_4^{2-}(\mathrm{aq})\)

The corresponding equilibrium constants for these reactions will be as follows.

⇒ \(K_{\mathrm{a}_1}=\frac{\left[\mathrm{H}_3 \mathrm{O}^{+}\right]\left[\mathrm{HSO}_4^{-}\right]}{\left[\mathrm{H}_2 \mathrm{SO}_4\right]} \text { and } K_{\mathrm{a}_2}=\frac{\left[\mathrm{H}_3 \mathrm{O}^{+}\right]\left[\mathrm{SO}_4^{2-}\right]}{\left[\mathrm{HSO}_4^{-}\right]} \text {. }\)

  • As you can see, the dibasic acid H2SO4 has two ionisation constants. Similarly, tribasic acids like phosphoric acid will have three ionisation constants.
  • Like polybasic acids, polyacidic bases also have the ionisation constants corresponding to the number of ionisation steps.
  • For example, a diacidic base like ethylenediamine has two ionisation constants \(K_{b_1} \text { and } K_{b_2}\). Table gives some values of the ionisation constants of polyprotic acids.
  • As you can seen in Table among all the ionisations in an individual polybasic acid the first ionisation constant is the largest followed by the second and then the third.
  • This is because after the first ionisation, we are left with a negatively charged ion and it is difficult to remove a proton from a negative ion.

Basic Chemistry Class 11 Chapter 7 Equilibrium Stepwise Dissociation Constants For Polyprotic Acids At 298K

Factors affecting acid strength: We have already seen how to compare the strengths of two acids. But why is one acid stronger than the other? Put simply, the acid strength is determined by the strength and polarity of the H—A bond.

  • The strength of the bond depends on the enthalpy change associated with the dissociation of the HA molecule into H and A.
  • The polarity of the H—A bond depends on the electronegativity of A and the ease with which electron transfer can occur from H to A resulting in H+ and A ions. The weaker and more polar the HA bond, the stronger is the acid.
  • Let us consider the acids of the halogen series HF, HCl, HBr and HI.
  • The variation in polarity is not very significant while we move from HF to Til but there is a considerable difference in the bond strength, which is the deciding factor here.

Basic Chemistry Class 11 Chapter 7 Equilibrium Factors Affecting The Acid And Bond Strength Increases

You already know that atomic size increases on moving down a group. Thus, the size increases from F to I and so the bond strength decreases, increasing the acidity from HF to HI.

  • This trend is generally true of acids in the same group of the periodic table. When we go down the group from O to Te, the acid strength increases as H2O<H2S<H2Se<H2Te.
  • For acids in the same row of the periodic table, the variation in bond strength is insignificant and the polarity of the bond is the deciding factor for acid strengths.
  • Let us consider the hydrides of a few elements in the second period—CH4, NH3, H2O and HF. As we move from C to F, electronegativity increases and hence the added strength also increases from CH4 to HF.
  • The electronegativities given are C, N, O and F respectively.

Basic Chemistry Class 11 Chapter 7 Equilibrium Factors Affecting Acid ANd Electronegativity Strength Increases

Oxoacids Oxoacids such as H2SO4, H2CO3 and HNO3 have an O—H bond and the acid strength is dependent on the strength of this bond. Any factor that weakens the O—H bond or increases its polarity, increases the strength of the acid.

  • In these acids, the nonmetallic atom is bonded on one side to the oxygen of O—H. The electronegativity of this atom and its oxidation number influences the acid strength.
  • In the case of oxoacids containing the same number of O—H groups and the same number of O atoms, as the electronegativity of the nonmetallic atom increases, the acid strength increases.
  • For example, consider HOCl, HOBr and HOI. The electronegativity decreases from Cl to I, and hence add strength increases from HOI to HOBr to HOCl.

Basic Chemistry Class 11 Chapter 7 Equilibrium Acid Strength Increases

With the increase in the electronegativity of the nonmetallic atom attached to the O—H bond, the bonded electron cloud is pulled more towards the atom, thus weakening the O—H bond, which then tends to break easily.

If the oxoacids contain the same nonmetallic atom but different numbers of oxygen atoms, then as the oxidation number of the atom increases, the acid strength increases. For example, among the oxoacids of chlorine, the strength of the acids increases as

Basic Chemistry Class 11 Chapter 7 Equilibrium Acid And Oxidation Strength Increases

Common-Ion effect in the ionisation of acids and bases: You are already aware how acetic acid ionises in water. You also know how to calculate the degree of dissociation of this acid. The equilibriun involved is \(\mathrm{CH}_3 \mathrm{COOH}+\mathrm{H}_2 \mathrm{O} \rightleftharpoons \mathrm{CH}_3 \mathrm{COO}^{-}+\mathrm{H}_3 \mathrm{O}^{+}\)

∴ \(K_{\mathrm{a}}=\frac{\left[\mathrm{CH}_3 \mathrm{COO}^{-}\right]\left[\mathrm{H}_3 \mathrm{O}^{+}\right]}{\left[\mathrm{CH}_3 \mathrm{COOH}\right]} .\)

If we add sodium acetate (the salt of the conjugate base of acetic acid) to the acetic acid solution, the hydrogen ion concentration changes (it decreases).

  • This is because the addition of acetate ions causes an increase in the concentration of the products and the equilibrium shifts to the left-hand side, reducing the dissociation of acetic acid.
  • This is according to Le Chatelier’s principle. This is called the common ion effect, which is simply the shift in equilibrium on the addition of a substance that provides more of an ion already involved in the equilibrium. Therefore, the concentration of H3O+ decreases.

Suppose acetate ions of concentration 0.1 M are added a 0.1-M acetic acid solution. Then the initial concentrations of the acetate ion and acetic acid solution are both 0.1 M.

Let the concentration of the acetic acid dissociated be x. Then the concentrations of the species present can be given as follows:

⇒ \(\mathrm{CH}_3 \mathrm{COOH}(\mathrm{aq})+\mathrm{H}_2 \mathrm{O} \rightleftharpoons \mathrm{CH}_3 \mathrm{COO}^{-}(\mathrm{aq})+\mathrm{H}_3 \mathrm{O}^{+}(\mathrm{aq})\)

Initial conc. \( 0.1 0.1 \sim 0\)

Change \(-x +x +x\)

Equilibrium conc. \(0.1-x 0.1+x x\)

⇒ \(K_{\mathrm{a}}=18 \times 10^{-5}=\frac{\left[\mathrm{CH}_3 \mathrm{COO}^{-}\right]\left[\mathrm{H}_3 \mathrm{O}^{+}\right]}{\left[\mathrm{CH}_3 \mathrm{COOH}\right]}\)

= \(\frac{(0.1+x)(x)}{0.1-x}\).

Since Ka is small for a very weak acid, x is small compared to 0.1 and we can make the approximation that 0.1 +x ≈ 0.1 ≈ 0.1-x.

∴ \(K_{\mathrm{a}}=18 \times 10^{-5}=\frac{x(0.1)}{0.1}=x\)

or x = [H3O+] = 1.8 x 10-5 M.

Example: Calculate the hydronium iott concentration of a 0.1-M solution of acetic acid, given Ka = 1.8 x 10-5.
Solution:

⇒ \(\mathrm{CH}_3 \mathrm{COOH}(\mathrm{aq})+\mathrm{H}_2 \mathrm{O}(\mathrm{l}) \rightleftharpoons \mathrm{CH}_3 \mathrm{COO}^{-}(\mathrm{aq})+\mathrm{H}_3 \mathrm{O}^{+}(\mathrm{aq})\)

⇒ \(K_{\mathrm{a}}=\frac{\left[\mathrm{CH}_3 \mathrm{COO}^{-}\left[\mathrm{H}_3 \mathrm{O}^{+}\right]\right.}{\left[\mathrm{CH}_3 \mathrm{COOH}\right]}\).

If x is the concentration of CH3COOH that dissociates, then the concentrations for the various species in the solution are as follows.

⇒ \(\mathrm{CH}_3 \mathrm{COOH}+\mathrm{H}_2 \mathrm{O} \rightleftharpoons \mathrm{CH}_3 \mathrm{COO}^{-}+\mathrm{H}_3 \mathrm{O}^{+}\)

Initial conc. \(0.1 0 0\)

Change \(-x +x +x\)

Equilibrium conc. \(0.1-x x x\)

Substituting the values at equilibrium in the expression for Ka,

⇒ \(K_{\mathrm{a}}=\frac{x \times x}{0.1-x}=\frac{x^2}{0.1-x}\)

This will result in a quadratic equation in x. However, it may be simplified by assuming that x << 0.1

Hence, \(K_{\mathrm{a}}=\frac{x^2}{0.1}\)

⇒ x2 = 0.1 x Ka = 0.1 x 1.8 x 10-5

x = 1.34 x 10-3

∴ [H3O+] = 1.34 x 10-3

Limitations of Bronsted-Lowry theory The Bronsted-Lowry theory could explain many reactions which the Arrhenius theory could not. However, it still could not account for certain acid-base reactions which do not involve the transfer of a proton.

1. Reactions between acidic oxides like CO2, SO2 and SO3 and basic oxides like CaO, BaO and MgO cannot be explained by the theory of proton transfer. Such reactions take place in the absence of a solvent.

⇒ \(\mathrm{CaO}+\mathrm{SO}_3 \longrightarrow \mathrm{CaSO}_4\)

⇒ \(\mathrm{MgO}+\mathrm{CO}_2 \longrightarrow \mathrm{MgCO}_3\)

2. The acidic behaviour of compounds like BF3 and AlCl3 cannot be explained by the Bronsted-Lowry theory.

Lewis theory: In 1923, G N Lewis, an American chemist, extended the concept of adds and bases further.

  • According to the Lewis theory, an acid is a substance which can accept a pair of electrons and a base is a substance which can donate a pair of electrons.
  • While this definition encompasses most add-base reactions, it has the drawback of being so general as to indude other types of reactions in its ambit.
  • According to the Lewis theory, an acid-base reaction involves the donation of a pair of electrons by a base to an add, leading to the formation of a coordinate bond between them.

Lewis bases Lewis bases can be of two types,

  1. Neutral molecules like H2O, RNH4, NH3 and ROH, in which
    one atom has at least one unshared pair of electrons, can act as Lewis bases,
  2. All negative ions (F, Cl, OH, CN, etc.) can behave as Lewis bases.

Lewis acids Species with vacant orbitals in the valence shell of one of the atoms can act as Lewis acids. Some examples of Lewis acids are as follows.

1. Simple cations like Ag+, Cu2+ and Fe3+ can accept a pair of electrons and act as Lewis acids.

2. Molecules which have an atom with an incomplete octet can behave as Lewis acids, for example, BF3 and AlCl3.

Basic Chemistry Class 11 Chapter 7 Equilibrium Lewis Acid And Base

3. Molecules in which the central atom has vacant orbitals may acquire more than an octet of valence electrons, for example, SnCl4, SiF4 and PF5, and can act as Lewis acids.

⇒ \(\underset{\substack{\text { Lewis acid } \\ \text { acid }}}{\mathrm{SiF}_4}+\underset{\text { Lewis base }}{2 \mathrm{~F}^{-}} \longrightarrow \mathrm{SiF}_6^{2-}\)

4. Molecules which have a multiple bond between two atoms of different electronegativities, for example, CO2 and SO2, also behave as Lewis acids.

In the CO2 (O=C=O) molecule, carbon being less electronegative than oxygen, acquires a slight positive charge (because one n electron pair shifts more towards oxygen) and can, thus, accept a pair of electrons.

⇒ \(\underset{\substack{\text { Lewis } \\ \text { acid }}}{\mathrm{CO}_2}+\underset{\substack{\text { Lewis } \\ \text { base }}}{\mathrm{OH}^{-}} \longrightarrow \mathrm{HCO}_3^{-}\)

  • While it is true that the definition of acids and bases proposed by Lewis is more general than the other definitions, it does not mean that it encompasses all acid-base reactions.
  • There are substances, for instance, which are Bronsted acids but not Leivis acids. Bronsted acids like HCl and H2SO4 can donate a proton but are not capable of accepting a pair of electrons, so they are not Lewis acids.

However, all Bronsted bases are also Lewis bases because a species capable of donating an electron pair (Lewis base) also has the tendency to accept a proton (Bronsted base).

⇒ \(\mathrm{NH}_3+\mathrm{H}_2 \mathrm{O} \longrightarrow \mathrm{NH}_4^{+}+\mathrm{OH}^{-}\)

Here NH3 is a Lewis base as well as n Bronsted base but H2O is a Bronsted acid, but not a Lewis add.

Limitations of Lewis theory The Lewis theory of adds and bases can explain the acidic and basic nature of many substances which cannot be explained by either the Bronsted-Lowry theory or the Arrhenius theory. Still it is not a flawless theory.

  1. This concept of adds and bases is so general that it labels all reactions leading to the formation of coordinate bonds as acid-base reactions.
  2. It does not explain the addic behaviour of adds like HCl and H2SO3 which do not form coordinate bonds with bases.
  3. The formation of coordination compounds is usually a slow process, so add-base reactions should be slow- according to the Lewis theory. But acid-base reactions are generally fast.
  4. The Lewis theory cannot be used to ascertain the relative strengths of adds and bases.
  5. The catalytic property of many adds is due to the H+(aq) ion. Since a Lewis add need not possess hydrogen, a Lewis acid need not have this property.

While dealing with acids and bases such as HCl, CH3COOH, NaOH and NH4OH in aqueous solutions, the Bronsted-LowTy concept is the most suitable and hence widely accepted.

Self Ionisation Of Water

So far we have considered water as a solvent and seen that it has both acidic and basic properties, i.e., it is amphoteric in nature. We will be using the Bronsted-Low’ry concept to understand the ionisation of water. In the presence of an add, it acts as a base while in the presence of a base, it acts as an acid.

Basic Chemistry Class 11 Chapter 7 Equilibrium Self Ionisation Of Water

In pure water, one molecule of water can donate a proton to another molecule of water in a reaction in which water acts as acid and base simultaneously. There are always some H3O+ and OH ions present in pure water.

or \(\mathrm{H}_2 \mathrm{O}+\mathrm{H}_2 \mathrm{O} \rightleftharpoons \mathrm{H}_3 \mathrm{O}^{+}+\mathrm{OH}^{-}\)

This is referred to as dissociation self-ionisation or auto-ionisation of water and is characterised by the equilibrium constant,

or, \(K_{\mathrm{c}}=\frac{\left[\mathrm{H}_3 \mathrm{O}^{+}\right]\left[\mathrm{OH}^{-}\right]}{\left[\mathrm{H}_2 \mathrm{O}\right]\left[\mathrm{H}_2 \mathrm{O}\right]}\)

Multiplying the denominator of the above expression ([H2O]2) by Kc (as H2O is in excess and is essentially constant), we get K=[H2O+][OH]

Kw is called the ionic product of water. It is a constant at a particular temperature. At 298 K its value is 1.0 x 1014 mol2 L-2. Such a small value of Kw indicates that the auto-ionisation of water does not occur to a large extent.

In fact, the reverse reaction, the formation of water by the reaction of H3O+ and OH essentially goes to completion as its equilibrium constant is

⇒ \(\frac{1}{1 \times 10^{-14}}=1 \times 10^{14}\)(As stated earlier the units have been dropped.)

Kw increases with temperature because the degree of ionisation of water increases with temperature.

In pure water both H3O+ and [OH] are equal.

⇒ \(\left[\mathrm{H}_3 \mathrm{O}^{+}\right]=\left[\mathrm{OH}^{-}\right]=\sqrt{1 \times 10^{-14}}=1 \times 10^{-7} \text { at } 298 \mathrm{~K} \text {. }\)

We can calculate the extent of dissociation of water molecules by finding the concentration of dissociated and undissociated water.

The concentration of dissociated water is simply the concentration of [H3O+] or [OH]

Therefore concentration of dissociated water =1 x 10-7 M at 298 K.

The molar concentration of pure water can be calculated from its density and volume.

Now, number of moles = \(\frac{\text { mass }}{\text { molar mass }}=\frac{D \times V}{\text { molar mass }}\) (since mass = density x volume)

Concentration of undissociated water =\(\left(\frac{1000 \mathrm{~g}}{\mathrm{~L}}\right)\left(\frac{1 \mathrm{~mol}}{18 \mathrm{~g}}\right)=55.4 \mathrm{~mol} \mathrm{~L}^{-1}\)

(The density of pure water is 1000 gL-1 and its molar mass is 18.0 g mol-1.)

Therefore degree of dissociation = \(\frac{\text { conc. of dissociated water }}{\text { conc. of undissociated water }}\)

= \(\frac{1 \times 10^{-7}}{55.4}=18 \times 10^{-9} \approx 2 \times 10^{-9} \text { or } \frac{2}{10^9} \text {. }\)

This means that out of 109 molecules of water, two molecules are ionised.

Aqueous solutions of acids and bases: The concentrations of H3O+ ions and OH ions are the same in pure water. When some acid or base is added to water, these concentrations do not remain equal, but the value of Kw remains the same (at a particular temperature).

  • If some acid is added to water, the concentration of H3O+ ions will increase.
  • Then according to Le Chatelier’s principle, the system will try to counteract the change and the reverse reaction will be favoured, i.e., H3O+ ions will combine with OH ions to form water molecules.
  • In other words, the equilibrium will shift backwards, so that the value of Kw may remain the same.

At the new equilibrium, the concentration of OH ions will be less than the concentration of H3O+ ions. The concentration of OHions can be obtained from the following relation.

⇒ \(\left[\mathrm{OH}^{-}\right]=\frac{K_{\mathrm{w}}}{\left[\mathrm{H}_3 \mathrm{O}^{+}\right]}\)

Similarly, the addition of a base will increase the concentration of OH ions and decrease the concentration of H3O+ ions.

The concentration of H3O+ ions in an aqueous solution of a base can be obtained from the following expression.

⇒ \(\left[\mathrm{H}_3 \mathrm{O}^{+}\right]=\frac{K_{\mathrm{w}}}{\left[\mathrm{OH}^{-}\right]} .\)

In pure water or neutral solutions [H3O+] = [OH].

In acidic solutions [H3O+]> [OH].

In basic solutions [H3O+]<[OH].

According to Arrhenius, a strong acid or a strong base completely dissociates into hydrogen ions and the corresponding anions or hydroxyl ions and the corresponding cations (as the case may be).

Thus, a 0.1-M HCl solution will dissociate to give 0.1-M H+ ions and 0.1-M Cl ions.

Example: Calculate the hydronium ion and hydroxyl ion concentrations in

  1. a 0.001-M HCl solution,
  2. a 0.01-M HNO3 solution,
  3. a 0.01-M NaOH solution,
  4. a 0.01-M Ba(OH)2 solution,
  5. a 0.1-M HCOOH which is 10% ionised and
  6. a solution containing 3.65 x 10-3 g of HCl per 100 mL.

Solution:

1. HCl is a strong acid which ionises completely in water, so[H3O+] = 0.00 1 M = 1 x 10-3 mol L-1.

The equilibrium can be represented as follows.

⇒ \(\mathrm{HCl}+\mathrm{H}_2 \mathrm{O} \rightleftharpoons \mathrm{H}_3 \mathrm{O}^{+}+\mathrm{Cl}^{-}\)

⇒ \(K_{\mathrm{w}}=\left[\mathrm{OH}^{-}\right]\left[\mathrm{H}_3 \mathrm{O}^{+}\right]\)

⇒ \({\left[\mathrm{OH}^{-}\right]=\frac{K_{\mathrm{w}}}{\left[\mathrm{H}_3 \mathrm{O}^{+}\right]}=\frac{1 \times 10^{-14}}{10^{-3}}=1 \times 10^{-11} \mathrm{~mol} \mathrm{~L}^{-1}}\)

2. The equilibrium in this case can be represented as \(\mathrm{HNO}_3+\mathrm{H}_2 \mathrm{O} \rightleftharpoons \mathrm{H}_3 \mathrm{O}^{+}+\mathrm{NO}_3^{-}\)

Being a strong acid HN03 too dissociates completely and

⇒ \({\left[\mathrm{H}_3 \mathrm{O}^{+}\right]=\left[\mathrm{HNO}_3\right]=0.01 \mathrm{M}=1 \times 10^{-2} \mathrm{~mol} \mathrm{~L}^{-1} . }\)

∴ \(\quad{\left[\mathrm{OH}^{-}\right]=\frac{K_{\mathrm{w}}}{\left[\mathrm{H}_0 \mathrm{O}^{+}\right.}=\frac{1 \times 10^{-14}}{10^{-2}}=1 \times 10^{-12} \mathrm{~mol} \mathrm{~L}^{-1}.}\)

3. Since NaOH is a strong base it ionises completely in water.

∴ \(\quad\left[\mathrm{OH}^{-}\right]=[\mathrm{NaOH}]=0.01 \mathrm{M}=10^{-2} \mathrm{~mol} \mathrm{~L}^{-1}\)

∴ \(\quad\left[\mathrm{H}_3 \mathrm{O}^{+}\right]=\frac{K_{\mathrm{w}}}{\left[\mathrm{OH}^{-}\right]}=\frac{1 \times 10^{-14}}{10^{-2}}=1 \times 10^{-12} \mathrm{~mol} \mathrm{~L}^{-1}\).

4. Since Ba(OH)2 is completely ionised and its acidity is 2, i.e.,1 mol of Ba(OH)2 gives 2 mol of OH ions,

⇒ \(\mathrm{Ba}(\mathrm{OH})_2 \longrightarrow \mathrm{Ba}^{2+}+2 \mathrm{OH}^{-}\)

⇒ \({\left[\mathrm{OH}^{-}\right]=2\left[\mathrm{Ba}(\mathrm{OH})_2\right]=2 \times 0.01=0.02 \mathrm{~mol} \mathrm{~L}^{-1} .}\)

∴ \({\left[\mathrm{H}_3 \mathrm{O}^{+}\right]=\frac{K_{\mathrm{w}}}{\left[\mathrm{OH}^{-}\right]}=\frac{1 \times 10^{-14}}{2 \times 10^{-2}}=5 \times 10^{-13} \mathrm{~mol} \mathrm{~L}^{-1} .}\)

5. HCOOH is a weak acid whose degree of dissociation (α) is 10% (given).

⇒ \(\underset{\mathrm{C}(1-\alpha)}{\mathrm{HCOOH}}+\mathrm{H}_2 \mathrm{O} \rightleftharpoons \underset{\mathrm{C} \alpha}{\mathrm{H}_3 \mathrm{O}^{+}}+\underset{\mathrm{C} \alpha}{\mathrm{HCOO}^{-}}\)

⇒ \({\left[\mathrm{H}_3 \mathrm{O}^{+}\right]=\mathrm{C} \alpha=0.1 \times 0.1=0.01=10^{-2} \mathrm{~mol} \mathrm{~L}^{-1} .}\)

⇒ \({\left[\mathrm{OH}^{-}\right]=\frac{K_{\mathrm{w}}}{\left[\mathrm{H}_3 \mathrm{O}^{+}\right]}=\frac{1 \times 10^{-14}}{10^{-2}}=10^{-12} \mathrm{~mol} \mathrm{~L}^{-1} .}\)

6. Molarity of a solution = no. of moles in1 L of solution.

The given solution contains 3.65 x 10-3 g of HCl in 100 mL or 0.1 L.

∴ M = \(\frac{3.65 \times 10^{-3}}{36.5 \times 0.1}\) [mol. wt. of HCl = 36.5] = 1 x 10-3

HCl is a strong acid.

∴ \({\left[\mathrm{H}_3 \mathrm{O}^{+}\right]=[\mathrm{HCl}]=1 \times 10^{-3} \mathrm{~mol} \mathrm{~L}^{-1} .}\)

∴ \({\left[\mathrm{OH}^{-}\right]=\frac{1 \times 10^{-14}}{10^{-3}}=1 \times 10^{-11} \mathrm{~mol} \mathrm{~L}^{-1} .}\)

The pH scale: From our discussions on the ionisation of water so far, it must have become clear that Kw (ionic product) remains constant and that for a neutral solution [H3O+] is equal to [OH], for an acidic solution [H3O+] > [OH] and for a basic solution [H3O+]< [OH].

  • Surely then if we know the H3O+ ion concentration in a solution we would have a fair idea about its relative acidic or basic character.
  • The hydrogen ion concentration may vary from, say, 1 M in a strong acid to 10-14 M in a strongly basic solution. Expressing H3O+ ion concentrations in this fashion would be rather cumbersome.
  • In 1909, P L Sorensen, a Danish chemist, proposed a more convenient way of expressing the
  • H3O+
  • ion concentration in a solution. The scale proposed by Sorensen is called the pH (power of hydrogen) scale, which may be defined as
  1. The magnitude of the negative power to which 10 must be raised to express the hydronium ion concentration; or
  2. The negative logarithm (base 10) of the hydronium ion concentration in mol L-1.
  • Say the hydronium ion concentration in a solution is 10-x Then its pH would be x, according to the first definition. This is simple enough and the first definition of pH seems to be apt.
  • But what if the H3O+ ion concentration in some solution is 1.5 x 10-4 M? The first definition would not be so convenient.
  • This is why the second definition is preferred. Actually, the second definition can be derived from the first.

Suppose [H3O+] = 10-x

Then log[H3O+] = log 10-x = -x log 10 = -x.

∴ x = pH = -log[H3O+].

If you happen to know the pH of a solution, you could easily calculate the concentration of H3O+ ions in it. [H3O+]= antilog (-pH).

The hydroxyl ion concentration in a solution can be expressed in terms of pOH.

pOH = – log[OH].

For any solution at 298 K

Kw = [H3O+][OH] = 10-14

∴ log[H3O+] + log[OH] = logKw = log 10-14 =-14

or -log[H3O+]-log[OH] = -logKw =14

or pH + pOH = pKw =14.

For a neutral solution, [H3O+] = [OH] = 10-7.

∴ pH = -log 10-7 =7.

This brings us to something very useful.

For an acidic solution, pH < 7.

For a basic solution, pH > 7.

For a neutral solution, pH = 7.

For practical purposes, the pH range is taken as 0 to 14, though theoretically, it may be possible to have H3O+ ion concentrations of more than 10° M (i.e., 1 M) or less than 10-14 M.

The table gives the pH values of some common substances in everyday life.

Basic Chemistry Class 11 Chapter 7 Equilibrium The pH Values Of Some Common Substances

Example 1. Calculate the pH value of a solution of

  1. 0.01-M HCl and
  2. 0.01-M NaOH.

Solution:

1. \(\mathrm{HCl}(\mathrm{aq})+\mathrm{H}_2 \mathrm{O}(\mathrm{l}) \longrightarrow \mathrm{H}_3 \mathrm{O}^{+}(\mathrm{aq})+\mathrm{Cl}^{-}(\mathrm{aq})\)

Since HCl is completely ionised,

[H3O+] = [HCl] = 0.01 M = 1 x 10-2 mol L-1.

pH = – log[H3O+] = – log 1 x 10-2 = – [-2] = 2.

2. \(\mathrm{NaOH}(\mathrm{aq}) \longrightarrow \mathrm{Na}^{+}(\mathrm{aq})+\mathrm{OH}^{-}(\mathrm{aq})\)

NaOH is a strong base which ionises completely.

[OH] = [NaOH] = 0.01 M = 10-2 mol L-1.

∴ \(K_{\mathrm{w}}=\left[\mathrm{H}_3 \mathrm{O}^{+}\right]\left[\mathrm{OH}^{-}\right]\)

∴  \(\quad {\left[\mathrm{H}_3 \mathrm{O}^{+}\right]=\frac{K_{\mathrm{w}}}{\left[\mathrm{OH}^{-}\right]}=\frac{1 \times 10^{-14}}{1 \times 10^{-2}}=10^{-12} \mathrm{~mol} \mathrm{~L}^{-1} . }\)

∴ \(\quad \mathrm{pH}=-\log \left[\mathrm{H}_3 \mathrm{O}^{+}\right]=-\log 10^{-12}=-[-12]=12\).

Example 2. 1000 mL of a solution contains 6.3 g of nitric acid. What is the pH of the solution if the acid is completely dissociated?
Solution:

The concentration of HNO3 = 6.3 g L-1.

Molecular weight of HNO3 = 63.

∴ M = 6.3/6.3 = 0.1.

⇒ \(\mathrm{HNO}_3+\mathrm{H}_2 \mathrm{O} \rightleftharpoons \mathrm{H}_3 \mathrm{O}^{+}+\mathrm{NO}_3^{-}\)

Since the acid is completely dissociated

[H3O+] = [HNO3]=0.1M = 10-1 M.

pH = – log[H3O+] = – log 10-1 = 1.

  • The pH scale is very useful in getting an idea about the relative acidic (or basic) strengths of solutions. The higher the pH, the lower the concentration of hydronium ions in the solution.
  • If the pH of a solution changes by one unit, there is a tenfold change in the molar concentration of that solution. Here it is important for you to know that pH is a measure of ‘hydrogen ion activity’.
  • At the present level of learning, for the sake of simplicity, we have been using hydrogen ion concentration as a measure of pH since we are considering ideal (dilute) behavior of solutions.
  • However, many solutions deviate from ideal behaviour with rise in concentration. To overcome this problem, we use the concept of activity.
  • Activity is a physical quantity, a thermodynamic function that can be used in place of concentration. Now consider a concentrated solution, for example, 1 M HCl.

Let us use both its activity and molar concentration to determine its pH.

⇒ \(a_{\mathrm{H}_3 \mathrm{O}^{+}} \text {(activity) }=\gamma\left[\mathrm{H}_3 \mathrm{O}^{+} \right.\)., where y is the activity coefficient, a dimensionless quantity.

The activity being 0.81 the pH value of the solution turns out to be 0.092. On the other hand, if we use the concentration, pH = – log[1] = 0.

The ‘p’ in pH indicates (-log) so that px means -log(x). This is very useful in calculations involving exponential numbers so that it has been extended to other species. The most widely used are pKw, pKa, and pKb.

PKa = -logKa; pKb =-logKb.

As you already know, for a conjugate acid-base pair, Ka x Kb = Kw.

Taking negative logarithm on both sides, (- log Ka) + (- log Kb ) = (- log Kw)

or PKa+pKb =pKw.

Measuring pH The pH of a solution can be measured with the help of pH papers, indicators, or pH meters. While pH papers and indicators give an approximate pH value of the solution in question, pH meters give the exact value.

  • An acid-base indicator is a substance that changes colour in a specific pH range of about 2 pH units. To determine the pH of a solution, a few drops of the indicator are added to the solution.
  • To make the determination easier, nowadays a universal indicator is available, which is simply a mixture of several indicators, to make approximate measurements in the pH range 3-10.
  • The color change on the addition of the indicator is noted and compared with a color chart to know the pH of the given solution.
  • Nowadays, pH meters are available for commercial purposes. They measure pH with high precision.

A typical pH meter consists of a special measuring probe (a glass electrode) connected to an electronic meter that measures and displays the pH reading.

Basic Chemistry Class 11 Chapter 7 Equilibrium A pH Meter With Its Probeimmersed In An Acidic Solution

Hydrolysis Of Salts

The term hydrolysis refers to any reaction in which water is one of the reactants. Water, being an amphiprotic solvent, can act both as an acid or a base. When a salt is dissolved in water, it dissociates into its ions. If the salt is represented as MA, then we have

⇒ \(\mathrm{MA}(\mathrm{aq}) \longrightarrow \mathrm{M}^{+}(\mathrm{aq})+\mathrm{A}^{-}(\mathrm{aq})\)

Now, water may interact with the ions in two ways:

⇒ \(\underset{\text { acid }}{\mathrm{M}^{+}(\mathrm{aq})}+\underset{\text { base }}{2 \mathrm{H}_2 \mathrm{O}(\mathrm{l})} \underset{\text { base }}{\mathrm{MOH}(\mathrm{aq})}+\underset{\text { adid }}{\mathrm{H}_3 \mathrm{O}^{+}(\mathrm{aq})}\) ….(1)

⇒ \(\underset{\text { base }}{\mathrm{A}^{-}(\mathrm{aq})}+\underset{\text { acid }}{\mathrm{H}_2 \mathrm{O}(\mathrm{l})} \underset{\text { add }}{\mathrm{HA}(\mathrm{aq})}+\underset{\text { base }}{\mathrm{OH}^{-}(\mathrm{aq})}\)…..(2)

As you can see in reaction (1), the base from which the cation is obtained is generated while in reaction (2), the acid from which the anion is obtained is generated. The salt MA is obtained by a neutralisation reaction between the base, MOH, and the acid, HA.

  • A closer look at the reverse reactions shows that M+ is the conjugate acid of the base MOH while A is the conjugate base of the acid HA.
  • On dissolving a particular salt in water, whether reaction (1) or (2) or both will occur and to what extent is dependent upon the strengths of the acid and the base from which the salt has been formed.
  • The strength of an acid is inversely proportional to that of its conjugate form. If HA is a strong acid, its conjugate base, A, will be weak and hence reaction (2) will occur only insignificantly.
  • Similarly, if the base MOH is strong, its conjugate acid M+ will be weak and reaction (1) will occur to a very small extent.
  • If both M+ and A are obtained from a strong base and a strong acid respectively, hydrolysis occurs to a negligible extent or we may say that it does not occur at all. One such example is NaCl.
  • Na+ is obtained from NaOH, a strong base, and Cl is obtained from HCl, a strong acid.
  • A solution of NaCl will not show any hydrolysis and the species present in solution will be only Na+ (aq), Cl(aq), and H2O. Hence the pH will be 7 and the solution is neutral.

When a salt dissociates into ions on dissolution in water, the ion which is the conjugate of a strong acid/base does not hydrolyse.

For example, sodium acetate, CH3COONa (formed from a strong base and a weak acid), will dissociate in water as \(\mathrm{CH}_3 \mathrm{COONa}(\mathrm{aq}) \longrightarrow \mathrm{CH}_3 \mathrm{COO}^{-}(\mathrm{aq})+\mathrm{Na}^{+}(\mathrm{aq}).\)

Here CH3COO is the conjugate base of the weak acid CH3COOH. Therefore it will hydrolyse forming OH ions in the solution.

⇒ \(\mathrm{CH}_3 \mathrm{COO}^{-}(\mathrm{aq})+\mathrm{H}_2 \mathrm{O}(\mathrm{l}) \rightleftharpoons \mathrm{CH}_3 \mathrm{COOH}(\mathrm{aq})+\mathrm{OH}^{-}(\mathrm{aq})\)….(3)

However, Na+ will not hydrolyse as it is obtained from a strong base, NaOH. Therefore, the solution will be basic due to the hydrolysis reaction (3), in which hydroxyl ions are produced.

Similarly, NH4Cl (formed from a strong acid and a weak base) dissociates as \(\mathrm{NH}_4 \mathrm{Cl}(\mathrm{aq}) \rightleftharpoons \mathrm{NH}_4^{+}(\mathrm{aq})+\mathrm{Cl}^{-}(\mathrm{aq})\)

In this case NH4+ is the conjugate acid of the weak base NH3 and hence undergoes hydrolysis as \(\underset{\text { acid }}{\mathrm{NH}_4^{+}(\mathrm{aq})}+\underset{\text { base }}{2 \mathrm{H}_2 \mathrm{O}(\mathrm{l})} \rightleftharpoons \underset{\text { base }}{\mathrm{NH}_4 \mathrm{OH}(\mathrm{aq})}+\underset{\text { acid }}{\mathrm{H}_3 \mathrm{O}^{+}(\mathrm{aq})}\)…(4)

Cl will not undergo hydrolysis as it is obtained from HCl (strong acid). As a result, a solution of NH4Cl will be acidic due to reaction (4), where H3O+ is produced.

  • Now let us see what happens when ammonium acetate, CH3COONH4, is dissolved in water.
  • Since the salt is formed from the weak acid CH3COOH and the weak base NH4OH, both the ions (CH3COO and NH4) formed will hydrolyse.
  • Since both OH and H3O+ are produced during hydrolysis, the pH of the solution will depend on the comparative strengths of the parent acid (Ka) and base (Kb) (from which the salt is formed).
  • If Ka < Kb, the solution will be alkaline as then the conjugate base of the acid (A) is stronger and will hydrolyse more as compared to the conjugate acid of the base (M+).

If Ka >Kb, the solution will be acidic. However, if Ka = Kb, the solution will be neutral. It is interesting to note that the pH of such solutions is independent of the concentration of the salt and is given by

pH=7 + 1/2(pKa-pKb)

Example 1. The pH of a 0.1-M solution of aniline, a weak base, is 8.8. Find its pKb.
Solution:

⇒ \(\underset{\text { anline }}(\mathrm{B})+\mathrm{H}_2 \mathrm{O} \rightleftharpoons \mathrm{BH}^{+}+\mathrm{OH}^{-}\)

Initial conc.(M) C 0 0

Change -x +x +x

Equilibrium cone. C-x x x

Since \(K_{\mathrm{b}}=\frac{\left[\mathrm{BH}^{+}\right]\left[\mathrm{OH}^{-}\right]}{[\mathrm{B}]}\),

∴ \(K_b=\frac{x^2}{C-x}\).

The pH of the solution = – log[H3O+] = 8.8.

∴ [H3O+] = antilog (-8.8) =16 x 10-9 M.

Since we know that \(\left[\mathrm{OH}^{-}\right]=\frac{K_w}{\left[\mathrm{H}_3 \mathrm{O}^{+}\right]},\)

⇒ \(\left[\mathrm{OH}^{-}\right]=\frac{1 \times 10^{-14}}{16 \times 10^{-9}}=6.25 \times 10^{-6}=x\)

Substituting the values of x and C in the expression for Kb we get

∴\(K_{\mathrm{b}}=\frac{\left(6.25 \times 10^{-6}\right)^2}{0.1-6.25 \times 10^{-6}}\)

Since 6.25 x 10-6 <<0.1, 0.1- 6.25 x 10-6 = 0.1

∴ \(K_{\mathrm{b}}=\frac{\left(6.25 \times 10^{-6}\right)^2}{0.1}=3.9 \times 10^{-10}\).

pKb =- log Kb =-log3.9 x 10-10

pKb =9.4

Example 2. The pKa of hypobromous acid (HOBr) is 8.7. Calculate the pH of a 0.1-M solution of the acid.
Solution:

⇒ \(\mathrm{HA}+\mathrm{H}_2 \mathrm{O} \rightleftharpoons \mathrm{H}_3 \mathrm{O}^{+}+\mathrm{A}^{-}\)

Equilibrium cone. (M) C- x x x

Since \(K_4=\frac{\left[\mathrm{H}_3 \mathrm{O}^{+}\right]\left[\mathrm{A}^{-}\right]}{[\mathrm{HA}]}=\frac{x^2}{C-x}\),

pKa =-log Ka =8.7.

Ka =antilog(-8.7) = 2 x 10-9

Now substituting the value of Ka and C in the expression for Ka, we get

2 x 10-9 = \(\frac{x^2}{0.1-x}\)

As Ka is small, 0.1 >> x.

∴ 2 x 10-9 = \(\frac{x^2}{0.1}.\)

or x2 =2×10-9– x 0.1 = 2 x 10-10

or x \(=\sqrt{2 \times 10^{-10}}=1.4 \times 10^{-5} \mathrm{M} .\)

This is the concentration of H3O+ ions

[H3O+]=x=1.4 X 10-5 M.

pH =- log[H3O+] = 4.8.

The pH of a 0.1-M solution of HOBr is 4.8.

Example 3. Calculate the degree of ionisation of 0.1-M acetic acid in the presence of 0.1-M HCl The pKa of acetic acid is 4.74. Compare it with its ionisation in pure water.
Solution:

⇒ \(\mathrm{HAc}+\mathrm{H}_2 \mathrm{O} \rightleftharpoons \mathrm{H}_3 \mathrm{O}^{+}+\mathrm{Ac}^{-}\)

⇒ \(K_{\mathrm{a}}=\frac{\left[\mathrm{H}_3 \mathrm{O}^{+}\right]\left[\mathrm{Ac}^{-}\right]}{[\mathrm{HAc}]}\)

or \(\mathrm{p} K_{\mathrm{a}}=-\log K_{\mathrm{a}}=4.74\)

or Ka = antilog(-4.74) =1.8 x 10-5.

Let us first calculate α in pure water.

If the equilibrium concentration of [H3O+] = Cα then that of Ac’ is also Cα and the concentration of acetic acid that remains undissociated is [HAc]= C(1 – α).

∴ \(K_4=\frac{(\mathrm{C} \alpha)(\mathrm{C} \alpha)}{C(1-\alpha)}\)

Since acetic add is a weak acid, α << 1

Hence \(K_{\mathrm{a}} \cong \frac{C^2 \alpha^2}{C}=C \alpha^2\)

or \(\alpha=\sqrt{\frac{K_a}{C}}=\sqrt{\frac{18 \times 10^{-5}}{0.1}}=0.013 .\)

Therefore the percentage dissociation is 1.3%. In the presence of 0.01-M HCl, the source of H3O+ is acetic acid as well as hydrochloric acid. The equilibrium concentrations are as shown below.

⇒ \(\begin{array}{ll}
\mathrm{HAc}+\mathrm{H}_2 \mathrm{O} & \rightleftharpoons \mathrm{H}_3 \mathrm{O}^{+}+\mathrm{Ac}^{-} \\
\mathrm{C}(1-\alpha) & \mathrm{C} \alpha+0.01 \quad \mathrm{C} \alpha
\end{array}\)

Now, \(K_{\mathrm{a}}=\frac{(C \alpha+0.01)(C \alpha)}{C(1-\alpha)}\)

We may neglect α in comparison to 1.

∴ \(\quad K_a \cong \frac{C^2 \alpha^2+0.01 C \alpha}{C} \)

or \(K_a=C \alpha^2+0.01 \alpha\)

or \(C \alpha^2+0.01 \alpha-K_a=0\).

This is a quadratic equation and a can be calculated using the formula

x = \(\frac{-b \pm \sqrt{b^2-4 a c}}{2 a}\)

∴\(\quad \alpha =\frac{-0.01 \pm \sqrt{(0.01)^2+4 \times C \times K_a}}{2 C}\).

Substituting for Ka and C, we get

α = \(\frac{-0.01 \pm \sqrt{(0.01)^2+4 \times 0.1 \times 18 \times 10^{-5}}}{2 \times 0.1}=\frac{-0.01 \pm 0.0104}{0.2} .\)

Discarding the negative root, a = 2 x 10-3 = 0.002

Percentage dissociation = 0.002 x 100 = 0.2%

This is far less than that in pure water. This shows that the equilibrium is affected in the presence of a common ion.

Buffer Solutions

Suppose you have a solution of a particular pH or a solution which has a certain concentration of H3O+ ions. If you add even a small amount of add or base to this solution, its pH will change—it will decrease if you add an add and increase if you add a base.

Both in nature and in industries, it is necessary to have solutions whose pH values do not change with the addition of a small amount of an add or a base.

The pH value of an ordinary solution may change even if it comes in contact with air. It can absorb carbon dioxide from the air and become more addic, for example.

  • A solution which can resist change in pH when a small amount of an acid or a base is added to it (or when it is diluted) is called a buffer solution. The ability of such a solution to resist change in pH is called buffer action.
  • A buffer solution can be acidic. Such a solution contains a weak acid and its salt with a strong base in equimolar quantities (a weak acid and a salt formed from its conjugate base), example, a solution of acetic acid and sodium acetate, which acts as a buffer solution around pH 4.75.
  • A basic buffer solution, on the other hand, contains equimolar quantities of a weak base and its salt with a strong acid (a weak base and a salt formed from its conjugate base), example, NH4OH and NH4Cl, which acts as a buffer around pH 9.25.
  • A buffer solution may also contain a single substance, generally the salt of a weak acid and a weak base, example, ammonium acetate (CH3COONH4).
  • In order to understand how a buffer solution maintains a constant pH value, let us consider a solution containing equimolar quantities of CH3COOH (weak acid) and CH3COONa (a salt of acetic acid with a strong base).

In an aqueous solution acetic acid will be weakly ionised and sodium acetate will be mostly dissociated.

⇒ \(\mathrm{CH}_3 \mathrm{COOH}(\mathrm{aq})+\mathrm{H}_2 \mathrm{O}(\mathrm{l}) \rightleftharpoons \mathrm{CH}_3 \mathrm{COO}^{-}(\mathrm{aq})+\mathrm{H}_3 \mathrm{O}^{+}(\mathrm{aq})\)

⇒ \(\mathrm{CH}_3 \mathrm{COONa}(\mathrm{aq}) \longrightarrow \mathrm{CH}_3 \mathrm{COO}^{-}(\mathrm{aq})+\mathrm{Na}^{+}(\mathrm{aq})\)

The solution will contain a large concentration of Na+ and CH3COO ions, a very small concentration of H3O+ ions, and a large amount of undissociated acetic acid molecules.

If a few drops of an acid are added to this solution, the H3O+ ions provided by the acid disturb the equilibrium (between ionised and un-ionised acid), and CH3COO ions in the solution combine with the H3O+ ions to counter the change.

⇒ \(\underset{\text { from buffer }}{\mathrm{CH}_3 \mathrm{COO}^{-}(\mathrm{aq})}+\underset{\text { from acid }}{\mathrm{H}_3 \mathrm{O}^{+}(\mathrm{aq})} \rightleftharpoons \mathrm{CH}_3 \mathrm{COOH}(\mathrm{aq})+\mathrm{H}_2 \mathrm{O}(\mathrm{l})\)

Thus, the H3O+ ions provided by the acid are neutralised by CH3COO ions from the buffer solution and there is no change in pH.

If a few drops of a base are added to the solution, the OH” ions from the base combine with the H3O+ ions of the solution to form water.

⇒ \(\mathrm{OH}^{-}(\mathrm{aq})+\mathrm{H}_3 \mathrm{O}^{+}(\mathrm{aq}) \longrightarrow 2 \mathrm{H}_2 \mathrm{O}(\mathrm{l})\)

  • This disturbs the equilibrium (by reducing the concentration of H3O+ ions) and in accordance with Le Chatelier’s principle, more CH3COOH molecules ionise, restoring the original concentration of H3O+ ions or the pH.
  • Blood is a natural buffer solution whose pH is about 7.4. The pH of sea water is also almost constant.
  • Buffer solutions are required for electroplating and in the manufacture of photographic material and dyes, among many other industrial processes.
  • The pH of culture media in biological laboratories also need to be resistant to change. The buffer solution of desired pH can be prepared by taking appropriate ratios of salt and acid or salt and base and by knowing the respective pKa and pKb values.
  • An equation which is used to calculate the pH of a buffer solution is as follows.

pH = \(\mathrm{p} K_{\mathrm{a}}+\log \frac{[\text { salt }]}{\text { [acid] }}\) (for acidic buffer)

or pH = \(\mathrm{p} K_{\mathrm{a}}+\log \frac{[\text { base }]}{[\text { salt }]}\)
(for basic buffer)

This is called the Henderson-Hasselbalch equation. In the equation [acid], [base], and [salt] denote the concentrations of the acid, base, and salt respectively used to prepare the buffer solution.

Note that, in both the equations, pKa appears and it may be calculated for the conjugate acid of the base by using the expression pKa =pKw -pKb.

Example 1. Calculate the pH of an acetic acid-sodium acetate buffer in which the concentrations of the acid and salt are 0.5 M and 0.25 M respectively. Ka = 1.8x 10-5 for acetic acid.
Solution:

Using the Henderson-Hasselbalch equation,

pH = \(\mathrm{pK}_{\mathrm{a}}+\log \frac{\text { [salt] }}{\text { [acid] }}\)

pH = \(-\log \left(18 \times 10^{-5}\right)+\log \frac{0.25}{0.5}\) = 4.74 – 0.3010 = 4.44.

Example 2. Find the pH of an ammonia-ammonium chloride buffer solution in which the concentrations are \(C_{\mathrm{NH}_3}\)=0.2M and \(C_{\mathrm{NH}_4^{+}}\)= 0.3 M. Given that Ka for NH3 = 1. 76 x 10-5.
Solution:

pKa + pKb=pKw

or pKa=pKw-pKb

pKa = 14 – (- log 1.76 x 10-5) = 14-4.75 = 9.25.

Using this value of pKa in the equation

pH = \(\mathrm{p} K_{\mathrm{a}}+\log \frac{[\text { base] }}{[\text { salt }]}\)

pH = \(9.25+\log \frac{[0.2]}{[0.3]}\) = 9.25-0.18 = 9.07.

Example 3. Determine the pH of a buffer solution containing 0.03-M boric acid and 0.043-M sodium borate, given that pKa for B(OH)3 =9.00.
Solution:

pH = \(p K_{\mathrm{a}}+\log \frac{\text { [salt] }}{\text { [acid] }}\)

= \(9+\log \frac{0.04}{0.03}\) = 9 + 0.12 = 9.12.

Solubility Equilibria

The solubility of different salts in water or any other solvent varies to a large extent. For instance, calcium chloride is hygroscopic (absorbs water vapour from the atmosphere) in nature, and, on the other hand, lithium fluoride is almost insoluble.

There are certain factors which influence the solubility of salts in solvents. Common salt dissolves readily in water because the energies of coulmbic or ionic interaction between the two ions of a salt are overcome due to the high dielectric constant of water.

  • Thus, in water Na+ and C- exist freely— the two do not attract each other. The dissolution of a salt in water has the effect of reducing the force between the ions, which then separate as a consequence.
  • Ion solvation (solvation enthalpy—energy released in the process of solvation) also affects tire solubility of a salt in water or any other solvent. An ion is solvated when surrounded by several molecules of solvent.
  • Solvation enthalpy is more for polar solvents and less for nonpolar solvents. Thus, salts do not dissolve in nonpolar solvents as the solvation enthalpy is not enough to overcome lattice enthalpy. Different salts exhibit different solubilities at different temperatures.
  • Salts are divided into three categories based on solubility. Salts of solubility 0.1 M or greater are considered soluble.
  • Salts of solubility between 0.1 M and 0.01 M are considered arc considered slightly soluble. And salts of solubility less than 0.01 M, though called insoluble, are actually sparingly soluble.
  • A solid dissolves in a solvent until the solution and the solid are in equilibrium. The solution, at equilibrium, is saturated and its molar concentration is the molar solubility of the solid.

Solubility product: BaSO4 and AgCl arc sparingly soluble in water. When such a sparingly soluble salt is mixed with water, a small
amount of it dissolves and makes the solution saturated.

  • The rest of it remains undissolved and an equilibrium is set up between the undissolved salt and the salt ions in solution.
  • At equilibrium, the rate of dissolution of ions from the undissolved solid is equal to the rate of precipitation of ions from the saturated solution.
  • You may have noticed that we are talking about an equilibrium between the undissolved solid and the ions in solution, and not the molecules in solution.

This is because the small amount of the salt that dissolves gets completely dissociated into ions. The equilibrium can be represented as follows.

⇒ \(\underset{\text { or electrolyte }}Undissolved salt \rightleftharpoons ions in solution\)

Let us consider solid AgCl in contact with its saturated aqueous solution. The dissolution of the salt may be represented as

⇒ \(\mathrm{AgCl}(\mathrm{s}) \rightleftharpoons \mathrm{Ag}^{+}(\mathrm{aq})+\mathrm{Cl}^{-}(\mathrm{aq})\)

As in the other cases we have considered before, the equilibrium constant would be K= \(\frac{\left[\mathrm{Ag}^{+}\right]\left[\mathrm{Cl}^{-}\right]}{[\mathrm{AgCl}]}\)

The concentration of the undissolved salt is constant (it can be taken as 1) irrespective of the amount of the solid salt present. (This is true for any pure solid.)

Then K[AgCl] = [Ag+][C] =Ksp

This constant (Ksp) is called the solubility product. In general, if a sparingly soluble salt Ax By dissociates in water, the equilibrium thus set up can be represented as

⇒ \(\mathrm{A}_x \mathrm{~B}_y \rightleftharpoons x \mathrm{~A}^{y+}+y \mathrm{~B}^{x-}\).

The solubility product may be expressed as Ksp = [Ay+]x[Bx-]y

The solubility product of a sparingly soluble salt (or electrolyte) at a given temperature may be defined as the product of the molar concentrations of its ions in a saturated solution, with each concentration term raised to the power equal to the number of ions of that species produced by the dissociation of one molecule of the electrolyte.

A few examples will make this clear.

⇒ \(\mathrm{Ca}_3\left(\mathrm{PO}_4\right)_2: K_{\text {sp }}  =\left[\mathrm{Ca}^{2+}\right]^3\left[\mathrm{PO}_4^{3-}\right]^2\)

⇒ \(\mathrm{BaSO}_4: K_{\text {sp }} =\left[\mathrm{Ba}^{2+}\right]\left[\mathrm{SO}_4^{2-}\right]\)

⇒ \(\mathrm{Mg}\left(\mathrm{OH}_2\right): K_{\text {sp }}=\left[\mathrm{Mg}^{2+}\right]\left[\mathrm{OH}^{-}\right]^2\)

⇒ \(\mathrm{Ag}_2 \mathrm{CrO}_4: K_{\text {sp }} =\left[\mathrm{Ag}^{+}\right]^2\left[\mathrm{CrO}_4^{2-}\right]\)

  • It is worthwhile to mention here that ionic product (or Qsp) is the product of the concentration of ions in any solution raised to their respective stoichiometries.
  • It may have struck you that both solubility product and ionic products are products of the concentrations of ions in a solution (not considering the powers of the concentration terms for the moment).
  • However, the ionic product has a more general application. It applies to both saturated and unsaturated solutions, while the solubility product applies only to saturated solutions.
  1. The ionic product is the same as the solubility product in a saturated solution.
  2. If the solubility product is greater than the ionic product, the solution is unsaturated.
  3. The higher the value of the solubility product the greater is the solubility of a salt.
  4. The ionic product cannot be greater than the solubility product. When it tends to exceed the value of the solubility product, precipitation (or combination of ions) starts.

If you know the solubility of a sparingly soluble salt at a particular temperature, you can easily calculate its solubility product.

Basic Chemistry Class 11 Chapter 7 Equilibrium Solubility Product Constants Of Some Compounds At 298K

Example 1. Find the solubility product of AgCl at T°C if its solubility at this temperature is 1.08 x 10-5 mol L-1.
Solution:

⇒ \(\mathrm{AgCl}(\mathrm{s}) \rightleftharpoons \mathrm{Ag}^{+}(\mathrm{aq})+\mathrm{Cl}^{-}(\mathrm{aq})\)

It is given that the solubility of AgCl is 1.08×10-5 mol L-1. AgCl is a sparingly soluble salt and it dissociates completely.

That is to say, 1 mol of AgCl in solution dissociates completely to produce 1 mol of Ag+ and 1 mol of Cl ions.

Therefore, 1.08 x 10-5 mol of AgCl will give 108 x 10-5 mol of Ag+ and 1.08 x 10-5 mol of Cl ions,

or [Ag+] = 1.08 x 10-5 mol L-1

and [Cl] = 1.08x 10-5 mol L-1

= (1.08 x 10-5 )(1.08 x 10-5) = 1.16 x 10-10.

Example 2. The solubility of PbCl2 at a particular temperature is 2.2 x 10-2 mol L-1. Find its solubility product at this temperature.
Solution:

⇒ \(\mathrm{PbCl}_2 \rightleftharpoons \mathrm{Pb}^{2+}+2 \mathrm{Cl}^{-}\)

PbCl2 is a sparingly soluble salt, so it is completely ionised in solution.

In other words, 1 mol of PbCl2 in solution will dissociate to produce 1 mol of Pb+ and 2 mol of Cl ions. It is given that the solubility of PbCl2 is 2.2 x 10-2 mol L-1.

[Pb2+] = 2.2 x 10-2 mol L-1

and [Cl] = 2 x 2.2 x 10-2 mol L-1.

∴ Ksp =[Pb2+][Cl]2 = [2.2 x 10-2][2 x 2.2 x 10-2]2 =4.3×10-5.

Calculation of solubility: The solubility of a salt is defined as grams of solute that can be dissolved in 100 mL of the solvent. The solubility of a sparingly soluble salt can be calculated if its solubility product is known.

  • However, since Ksp is obtained considering the concentration in moles per litre (M), solubility is also obtained in moles per litre.
  • This is called molar solubility—the amount of n salt that dissolves in a requisite amount of water to produce one litre of a saturated solution.
  • The following examples illustrate the calculation of molar solubility from solubility product.

Example 1. The solubility product of Ag2CrO4 at 298 K is 4 x 10-12. Find its solubility at this temperature.
Solution:

At equilibrium \(\mathrm{Ag}_2 \mathrm{CrO}_4 \rightleftharpoons 2 \mathrm{Ag}^{+}+\mathrm{CrO}_4^{2-}\)

Suppose the solubility of Ag2CrO4 is x mol L-1

Then [Ag+]=2x and [CrO42-] = x. (Ag2CrO4 forms 2 moles of Ag+ and one mole of CrO42-)

∴ \(K_{\mathrm{sp}}=\left[\mathrm{Ag}^{+}\right]^2\left[\mathrm{CrO}_4^{2-}\right]=[2 x]^2[x]=4 x^3\)

Given that Ksp = 4 x 10-12

or 4x3 =4 x 10-12

∴ x = 1 x 10-4

Therefore the solubility of Ag2CrO4 is 1 x 10-4 mol L-1.

Example 2. The solubility product of AgBr in water is 2.5 x 10-13. Calculate its solubility in a 0.01-M NaBr solution.
Solution:

NaBr dissociates completely in solution.

∴ \(\left[\mathrm{Br}^{-}\right]=[\mathrm{NaBr}]_0=0.01 \mathrm{M}\) (The subscript ‘0’ denotes the initial concentration.)

Let the solubility of AgBr be x mol L-1

Then [Ag+] = [Br] = x mol L-1 (from the dissociation of AgBr)

∴ total[Br] = x + 0.01 M (x <<0.01 M; AgBr being sparingly soluble)

Ksp(AgBr) = [Ag+][Br] = x x 0.01

Given that Ksp(AgBr) = 2.5 x 10-13

∴ x x 0.01 = 2.5 x 10-13

or x = \(\frac{2.5 \times 10^{-13}}{0.01}=2.5 \times 10^{-11} \mathrm{~mol} \mathrm{~L}^{-1}\)

Factors affecting solubility: The principle that an equilibrium counteracts the change due to the addition of species to solutions is also applicable to solubility product constants.

Common-ion effect: The presence of a common ion can also affect the solubility equilibrium. Let us consider the dissolution of MgF2 in water.

⇒ \(\mathrm{MgF}_2(\mathrm{~s}) \rightleftharpoons \mathrm{Mg}^{2+}(\mathrm{aq})+2 \mathrm{~F}^{-}(\mathrm{aq})\)

Let x be the solubility of MgF2 in moles per litre. Then the concentrations of Mg2+ and F are as follows:

[Mg2+] = x mol L-1 (as one mole of MgF2 gives one mole of Mg2+ and 2 moles of F).

[F] = 2 xmol L-1.

∴ Ksp =[Mg2+][F]2 = x x(2x)2 =4x3.

But the solubility product of MgF2 is 7.4 x 10-11

∴ 7.4 x 10-11 = 4x3

or \(x^3=\frac{7.4 \times 10^{-11}}{4}=1.8 \times 10^{-11} \Rightarrow x=\sqrt[3]{1.8 \times 10^{-11}}=2.6 \times 10^{-4} \mathrm{~mol} \mathrm{~L}^{-1}\)

∴ \({\left[\mathrm{Mg}^{2+}\right]=2.6 \times 10^{-4} \mathrm{~mol} \mathrm{~L}^{-1} .}\)

Molar solubility = 2.6 x 10-4 M.

∴ [F] = 2 x 2.6 x 10-4 =5.2×10-4M.

When NaF is added to a solution of MgF2, F being the common ion, its concentration in solution increases.

In order to keep Ksp constant, [Mg2+] must become smaller, or in other words, MgF2 is less soluble in an NaF solution than it is in pure water.

The presence of a common ion shifts the equilibrium to the left. This may be seen clearly from the following solved example.

Example: Calculate the molar solubility of MgF2 in 0.05-M NaF at 298 K, given that Kspof MgF2 is 7.4 x 10-11.
Solution:

We have calculated the molar solubility of MgF2 in pure water, which is 2.6×10-4 M. Let us now calculate its solubility in NaF.

If we assume x to be the molar solubility of MgF2 then one mole of MgF2 gives 1 mol of Mg2+ ions and 2 moles of F ions.

Hence the concentrations of Mg2+ and F from MgF2 alone are x and 2x respectively. There is another source of F, which is NaF. From 0.05 M of NaF, we get 0.05 M of F ions.

Equilibrium conc. \(\mathrm{MgF}_2(\mathrm{~s}) \rightleftharpoons \underset{x}{\mathrm{Mg}^{2+}(\mathrm{aq})}+\underset{2 x+0.05}{2 \mathrm{~F}^{-}(\mathrm{aq})}\)

Hence [F] = 2x+ 0.05 and [Mg2+] = x.

But Ksp =[Mg2+][F]2 or 7.4 x 10-11 = x x (2x + 0.05)2,

As Ksp is smqll 2x <<0.05.

Hence 7.4 x 10-11 s x x (0.05)2 or x = 7.4 x 10-9 M.

The molar solubility of MgF2 has decreased from 2.6 x 10-4 M in pure water to 7.4 x 10-9 M in the presence of NaF due to the common-ion effect.

pH of a solution: The solubility of salts of weak acids increases as the pH of the solution decreases or the acidity increases.

For example, the solubility of CaCO3 increases with decreasing pH.

This is because the CO32- ions react with H+ ions forming HCO3 ions react with H and thus are removed from the solution. As a result the reaction moves in the forward direction, dissolving more CaC03.

⇒ \(\begin{aligned}
\mathrm{CaCO}_3(\mathrm{~s}) \rightleftharpoons \mathrm{Ca}^{2+}(\mathrm{aq})+\mathrm{CO}_3^{2-}(\mathrm{aq}) \\
\mathrm{H}_3 \mathrm{O}^{+}(\mathrm{aq})+\mathrm{CO}_3^{2-}(\mathrm{aq}) \rightleftharpoons \mathrm{HCO}_3^{-}(\mathrm{aq})+\mathrm{H}_2 \mathrm{O}(\mathrm{l}) \\
\hline
\mathrm{CaCO}_3(\mathrm{~s})+\mathrm{H}_3 \mathrm{O}^{+}(\mathrm{aq}) \rightleftharpoons \mathrm{Ca}^{2+}(\mathrm{aq})+\mathrm{HCO}_3^{-}(\mathrm{aq})+\mathrm{H}_2 \mathrm{O}(\mathrm{l}) \\
\hline
\end{aligned}\)

Salts containing anions such as \(\mathrm{PO}_4^{3-}, \mathrm{S}^{2-}, \mathrm{F}^{-} \text {and } \mathrm{CN}^{-}\) also behave in a similar fashion.

Predicting the precipitation of a salt: If the value of the ionic product exceeds that of the solubility product, precipitation of the excess ions occurs.

So, if you know the solubility product of a sparingly soluble salt, you will be able to predict whether or not mixing two solutions containing known concentrations of Us Ions will result in the precipitation of the salt.

Example 1. Predict whether precipitation will occur when equal volumes of a 0.02-M Na2SO4 solution and a 0.02-M BaC2 solution are mixed.(Ksp of BaSO4 = 1.5 x 10-10)
Solution:

Let the volume of each solution = V mL

Total volume after mixing = 2V mL

Applying the relation \(\begin{aligned}
M_1 V_1=M_2 V_2 \\
0.02 \times V=M_2 \times 2 V
\end{aligned}\)

∴ after mixing [BaCl2]0 = 0.01 mol L-1

Since BaCl2, is completely ionised,

[Ba2+] =[BaCl2]0 = 0.01 mol L-1

To determine the molarity of the Na2SO4 solution after mixing, we use the equation

M’1V’1(before mixing) = M’2V’2(after mixing)

∴ \(M_2^{\prime}=\frac{0.02 \times V}{2 V}=0.01\)

∴ after mixing[Na2SO4]0 = 0.01 mol L-1

Since Na2SO4is completely ionised,

[SO42-] = [Na2SO4] = 0.01 mol L-1

Ionic product of BaSO4, LP. or Qsp = [Ba2+][SO42-] = 0.01 x 0.01 = 10-4 mol L-1

As Qsp >  Ksp precipitation occurs.

Precipitation of soluble salts: The precipitation of soluble salts from their saturated solutions is called salting out This process is used to obtain pure sodium chloride from a saturated solution of impure sodium chloride.

  • HCl gas is passed through a saturated solution of NaCL The addition of HCl (which ionises) increases the concentration of Cl ions.
  • Consequently, the ink product exceeds the solubility product of NaCl and precipitation occurs. The salt which is precipitated is in the pure stale. The impurities remain in solution.
  • Soap is salted out from its saturated solution by the addition of sodium chloride.
  • Soaps are sodium salts of ratty acids. The addition of sodium chloride increases the concentration of Na+ ions in the solution, thus making the ionic product exceed the solubility product

Example 1. Calculate the solubility of BaSO4 in pure water and in sea -water which contains 0.029-M SO42- ions. [Ksp(BaSO4) = 9.1x 10-11n. ]
Solution:

⇒ \(\mathrm{BaSO}_4(\mathrm{~s}) \rightleftharpoons \mathrm{Ba}^{2+}(\mathrm{aq})+\mathrm{SO}_4^{2-}(\mathrm{aq})\)

Let x be the molar solubility of BaSO4.

∴ \(\left[\mathrm{Ba}^{2-}\right]=x ;\left[\mathrm{SO}_4^{2-}\right]=x\)

⇒ [latexK_{\mathrm{sp}}=\left[\mathrm{Ba}^{2+}\right]\left[\mathrm{SO}_4^{2-}\right]=x^2 .[/latex]

But \(9.1 \times 10^{-11}=x^2 \text { or } x=\sqrt{9.1 \times 10^{-11}}=9.5 \times 10^{-6}\) .

Hence the solubility of BaSO4 in pure water = 9.5 x 10-6 M

In sea water, [SO42-] = 0.029 M.

Total concentration of SO42- = 0.029 + x’, where x’ is the solubility of BaSO4 in sea water.

Now, Ksp = (x’)(0.029 + x’)

We may assume that x'<< 0.029 as Ksp is small.

∴ Ksp =(x’)(0.029)

∴or 9.1 x 10-11 = 0.29x’

or x’ = 3.1x 10-10

Therefore, the solubility of BaSO4 in sea water is 3.1 x 10-10 M. This is far less than that in pure water.

Example 2. A carbonated drink was saturated with CO2 at 0°C and 5 atm pressure. What was the concentration of CO2 in the bottled drink if the Henry constant for an aqueous solution of CO2 at 0°C is 7.5 x 10-2 mol L-1 atm-1?
Solution:

According to Henry’s law, C = kp,

where C is the concentration of a gas in a solution, p is the partial pressure of the gas and k is the Henry constant.

∴ concentration of CO2 in the drink = kp = 7.5 x 10-2 x5

=37.5 x 10-2 =0.38 mol L-1.

Example 3. Calculate the equilibrium constantfor a reaction given that the rate constantsfor thefonvard and reverse reactions are 2.48 x 10-4 and 8.25 x10-5 respectively.
Solution:

Equilibrium constant K = \(\frac{k}{k^{\prime}}\),

where k and k’ are the rate constants for the forward and the reverse reaction respectively.

∴ K = \(=\frac{2.48 \times 10^{-4}}{8.25 \times 10^{-5}}=3\)

Example 4. Two moles of PCl5 was heated to 327°C in a closed vessel of volume 2 L. When equilibrium was established, 40% of the PCl5 had dissociated into PCI3 and Cl2 Calculate the equilibrium constantfor the reaction.
Solution:

The dissociation of PCl5 into PCI3 and Cl2 can be represented as follows.

⇒ \(\mathrm{PCl}_5 \rightleftharpoons \mathrm{PCl}_3+\mathrm{Cl}_2\)

Initial concentration of PCl5 = 2 mol.

Percentage dissociation at equilibrium = 40%.

PCI5 dissociated at equilibrium = 40% of 2 mol = 0.8 mol.

∴ amount of PCI5 at equilibrium = 2- 0.8 =12 mol

and amount of PCI3 and Cl2 at equilibrium = 0.8 mol.

Since the volume of the vessel is 2 L, at equilibrium

⇒ \(\left[\mathrm{PCl}_5\right]=\frac{1.2}{2} \mathrm{~mol} \mathrm{~L}^{-1}\)

⇒ \(\left[\mathrm{PCl}_3\right]=\frac{0.8}{2} \mathrm{~mol} \mathrm{~L}^{-1}\).

⇒ \(\left[\mathrm{Cl}_2\right]=\frac{0.8}{2} \mathrm{~mol} \mathrm{~L}^{-1}\).

∴ K=\(\frac{\left[\mathrm{PCl}_3\right]\left[\mathrm{Cl}_2\right]}{\left[\mathrm{PCl}_5\right]}=\frac{0.4 \times 0.4}{0.6}=0.267\).

Example 5. 0.1 mol of PCl5 is heated in a 1-L flask at 250°C. Calculate the equilibrium concentrations of PCl5, PCl3 and Cl2, if the equilibrium constant for the dissociation of PCl5 is 0.0414.
Solution:

PCI5 dissociates according to the following reaction.

∴ \(\mathrm{PCl}_5 \rightleftharpoons \mathrm{PCl}_3+\mathrm{Cl}_2\)

Suppose at equilibrium x mol of PCl5 dissociates to form x mol of PCl3 and x mol of Cl2. Since the volume of the vessel is 1 L, at equilibrium,

⇒ [PCl5] = 0.1 – x mol L-1.

⇒ [PCl3] = xmolL

⇒ [Cl2] = x mol L-1.

∴ K = \(\frac{\left[\mathrm{PCl}_3\right]\left[\mathrm{Cl}_2\right]}{\left[\mathrm{PCl}_5\right]}=\frac{x \times x}{0.1-x}\)

Given that K = 0.0414.

∴ 0.0414 = \(\frac{x^2}{0.1-x}\)

or x2 +0.0414x-0.00414 = 0

This is a quadratic equation and x can be calculated using the formula

x =\(-\frac{b \pm \sqrt{b^2-4 a c}}{2 a}=\frac{-0.0414 \pm \sqrt{(0.0414)^2-4 \times 1 \times(-0.0414)}}{2 \times 1}\)

=\(\frac{-0.0414 \pm \sqrt{0.0017+0.1656}}{2}=\frac{-0.0414 \pm 0.135}{2}\) = 0.0882 and + 0.0468

The negative value of x is meaningless.

Thus, the concentrations of PCI3, Cl2 and PCl5 at equilibrium are as follows.

[PCI3] = [Cl2] = x = 0.0468 mol LT-1.

[PCI5] = 0.1 – x = 0.0532 mol L-1.

Example 6. The equilibrium constant for the reaction \(\mathrm{CH}_3 \mathrm{COOH}+\mathrm{C}_2 \mathrm{H}_5 \mathrm{OH} \rightleftharpoons \mathrm{CH}_3 \mathrm{COOC}_2 \mathrm{H}_5+\mathrm{H}_2 \mathrm{O}\) is 4.0 at 298 K. Calculate the weight of ethyl acetate that will be obtained when 120 g of acetic acid reacts with 92 g of alcohol.
Solution:

Weight of ethyl alcohol = 92 g.

Weight of acetic acid = 120 g.

Number of moles = \(\frac{\text { weight of substance }}{\text { molecular weight }} \text {. }\)

∴ number of moles of ethyl alcohol = 92/46 = 2

and number of moles of acetic acid = 120/60 = 2.

According to the chemical equation that represents the reaction, 1 mol each of the reactants produce 1 mol each of the products. Therefore, if x mol of each of the reactants reacted, x mol each of the products would be formed at equilibrium.

⇒ \(\mathrm{CH}_3 \mathrm{COOH}+\mathrm{C}_2 \mathrm{H}_5 \mathrm{OH} \rightleftharpoons \mathrm{CH}_3 \mathrm{COOC}_2 \mathrm{H}_5+\mathrm{H}_2 \mathrm{O}\)

Intial concentration \(\frac{2}{V} \frac{2}{V} \frac{0}{V} \frac{U}{V}\)

Conc. at equilibrium \(\frac{2-x}{V} \frac{2-x}{V} \frac{x}{V} \frac{x}{V}\)

Applying the law of equilibrium

K = \(\frac{\left[\mathrm{CH}_3 \mathrm{COOC}_2 \mathrm{H}_5\right]\left[\mathrm{H}_2 \mathrm{O}\right]}{\left[\mathrm{CH}_3 \mathrm{COOH}\right]\left[\mathrm{C}_2 \mathrm{H}_5 \mathrm{OH}\right]}\)

= \(\frac{[x / V][x / V]}{\left[\frac{2-x}{V}\right]\left[\frac{2-x}{V}\right]}=\frac{x^2}{(2-x)^2}\).

Given that K = 4.

∴ 4 = \(\frac{x^2}{(2-x)^2}\)

or 2 = \(\frac{x}{2-x}\)

or 4-2x = x

or x = 4/3 =1.33 mol.

Hence the amoimt of ethyl acetate formed = 1.33 mol

=1.33 x molecular weight of ethyl acetate

=1.33 x 88 =117.04 g

Example 7. The equilibrium constant at 298 K for \(\mathrm{Cu}(\mathrm{s})+2 \mathrm{Ag}^{+}(\mathrm{aq}) \rightleftharpoons \mathrm{Cu}^{2+}(\mathrm{aq})+2 \mathrm{Ag}(\mathrm{s})\) is 4.0 x 1015. In a solution in which copper has displaced some silver ions, the concentration of Cu2+ ions is 16 x 10-2 mol L-1 and the concentration of Ag+ ions is 2.0 x 10-9 mol L-1. Is this system at equilibrium?
Solution:

Applying the law of chemical equilibrium, assuming that the system is at equilibrium,

⇒ \(K_{\mathrm{c}}=\frac{\left[\mathrm{Cu}^{2+}(\mathrm{aq})\right][\mathrm{Ag}(\mathrm{s})]^2}{\left[\mathrm{Ag}^{+}(\mathrm{aq})\right]^2[\mathrm{Cu}(\mathrm{s})]} .\)

By convention, concentration of a solid = 1.

∴\(K_{\mathrm{c}}=\frac{\left[\mathrm{Cu}^{2+}(\mathrm{aq})\right]}{\left[\mathrm{Ag}^{+}(\mathrm{aq})\right]^2}\)

Substituting the concentrations of Cu2+ and Ag+ ions

∴ \(K_c=\frac{1.6 \times 10^{-2}}{\left(2 \times 10^{-9}\right)^2}=4 \times 10^{15}\)

This is the value of Kc for the reaction in equilibrium. Hence the given system is in equilibrium.

Example 8. The equilibrium constant for the reaction \(\mathrm{H}_2(\mathrm{~g})+\mathrm{I}_2(\mathrm{~g}) \rightleftharpoons 2 \mathrm{HI}(\mathrm{g})\) is 50.5 at 718 K. Predict the direction in which the reaction will proceed to reach equilibrium at 718 K if we start with 2 x 10-2 mol of HI,1.0 x 10-2 mol of H2 and 3.0 x 10-2 mol of I2 in a vessel of volume1 L

Solution:

Reaction quotient \(\left(\mathrm{Q}_c\right)=\frac{[\mathrm{HI}]^2}{\left[\mathrm{H}_2\right]\left[\mathrm{I}_2\right]}\)

If the reaction quotient is equal to Kc, the system is at equilibrium, and if it is less than Kc, the reaction will proceed in the forward direction to attain equilibrium.

Since the volume of the vessel is 1 L,

⇒ \(Q_c=\frac{\left[2 \times 10^{-2}\right]^2}{\left[1.0 \times 10^{-2}\right]\left[3.0 \times 10^{-2}\right]}=1.33\)

Since this is less than Kc, the reaction will proceed in the forward direction.

Example 9. 1 mol of ammonia was injected into a 1-L flask at a certain temperature. The equilibrium mixture was found to contain 0.3 mol of H2. Calculate the concentrations of N2 and NH3 at equilibrium, and the equilibrium constant.
Solution:

The equilibrium can be represented as follows.

⇒ \(2 \mathrm{NH}_3 \rightleftharpoons \mathrm{N}_2+3 \mathrm{H}_2\)

The initial concentration of NH3 =1 mol.

At equilibrium, 2 mol of ammonia produces 3 mol of hydrogen.

Therefore, 0.3 mol of H2 will be obtained from 0.2 mol of NH3.

Similarly, when 2 mol of NH3 dissociates, 1 mol of N2 is obtained.

Therefore, 0.2 mol of NH3 will produce 0.1 mol of N2.

Since the volume of the flask is1 L, at equilibrium,

[N2] = 0.1 mol -1,

[H2] = 0.3 mol L-1,

[NH3] =1-0.2 =0.8 mol L-1.

∴ \(K_{\mathrm{c}}=\frac{\left[\mathrm{N}_2\right]\left[\mathrm{H}_2\right]^3}{\left[\mathrm{NH}_3\right]^2}=\frac{(0.1)(0.3)^3}{(0.8)^2}=0.004\)

Example 10. 1 mol of ammonium carbamate dissociates as shown below at 500 K. \(\mathrm{NH}_2 \cdot \mathrm{COONH}_4(\mathrm{~s}) \rightleftharpoons 2 \mathrm{NH}_3+\mathrm{CO}_2\) If the pressure exerted by the released gases is 3.0 bar, calculate the value of Kp.
Solution:

Applying the law of equilibrium,

⇒ \(K_p=\frac{p_{\mathrm{NH}_3}^2 p_{\mathrm{CO}_2}}{p_{\mathrm{NH}_2 \cdot \mathrm{COONH}_4(s)}}\)

By convention  \(p_{\mathrm{NH}_2 \mathrm{COONH}}\) =1 as it is a solid.

Given that the pressure exerted by the gases = 3 bar; this is the total pressure.

(To obtain the partial pressures, the total pressure must be multiplied by the mole fractions of the respective gases.)

∴ \(p_{\mathrm{NH}_3}=3 \times \frac{2}{3}=2 \mathrm{bar}\).

and \(p_{\mathrm{CO}_3}=3 \times \frac{1}{3}=1 \mathrm{bar}\).

∴ \(K_p=(2)^2(1)=4 bar.\)

Example 11. The dissociation constant of a weak acid HA is 1.6 x 10-5 at 298 K. Calculate the H3O+ ion concentration in a 0.01-M solution of the acid.
Solution:

The ionisation of the acid may be represented as

⇒ \(\mathrm{HA}(\mathrm{aq})+\mathrm{H}_2 \mathrm{O}(\mathrm{l}) \rightleftharpoons \mathrm{H}_3 \mathrm{O}^{+}(\mathrm{aq})+\mathrm{A}^{-}(\mathrm{aq})\)

⇒ \(K_a=\frac{\left[\mathrm{H}_3 \mathrm{O}^{+}\right]\left[\mathrm{A}^{-}\right]}{[\mathrm{HA}]}\) .

If the concentration of HA that dissociates is x, then the concentration of H3O+ and A is also x.

Then the concentration of the undissociated acid is (0.01- x).

Since HA is a weak acid, its degree of dissociation is small and at equilibrium, x may be neglected in comparison to 0.1.

[HA] = (0.1 -x) = 0.1 mol L-1

[H3O+] =[A] = X

∴ \(\quad K_a=\frac{x^2}{0.1}\)

or \(x^2=K_a \times 0.1=16 \times 10^{-5} \times 0.1\)

or \(x=\sqrt{16 \times 10^{-5} \times 0.1}=126 \times 10^{-3}\).

∴ \(\quad\left[\mathrm{H}_3 \mathrm{O}^{+}\right]=\left[\mathrm{A}^{-}\right]=1.26 \times 10^{-3} \mathrm{~mol} \mathrm{~L}^{-1}\).

Example 12. Calculate the degree of dissociation and the concentration of H3O+ ions in a 0.1-M solution of formic acid, given that =2.1 x 10-4 at 298 K.
Solution:

Formic acid dissociates in water according to the following equation.

⇒ \(\mathrm{HCOOH}+\mathrm{H}_2 \mathrm{O} \rightleftharpoons \mathrm{H}_3 \mathrm{O}^{+}+\mathrm{HCOO}^{-}\)

The initial concentration of the acid is given as 0.1 mol L-1

Let the degree of dissociation be α.

Then at equilibrium

[H3O+] = α x 0.1 mol L-1

[HCOO] =α x 0.1 mol L-1

[HCOOH] = 0.1(1- α) mol -1

∴\(K_a=\frac{\left[\mathrm{H}_3 \mathrm{O}^{+}\right]\left[\mathrm{HCOO}^{-}\right]}{[\mathrm{HCOOH}]}=\frac{(0.1 \alpha)(0.1 \alpha)}{0.1(1-\alpha)}\)

Since formic add is a weak acid, a is very small as compared to 1 and can be neglected in the denominator.

∴ \(K_a=\frac{(0.01 \alpha)^2}{0.1}=0.1 \times \alpha^2\)

Given that Ka = 2.1 x 10-4.

or 2.1 x 10 = 0.1 x α2

or \(\frac{21 \times 10^{-4}}{0.1}=\alpha^2\)

or α2 = 2.1 x 10-3 = 21 x 10-4

or α = 0.046

[H3O+] = Cα = 0.1 x 0.046 = 4.6 x 10-3 mol L-1

Example 13. A solution of NaOH contains 4.0 g of the base per litre. Find the pH of the solution.
Solution:

Concentration of NaOH solution = 4 g L-1; molecular weight of NaOH = 40.

∴ molarity of solution = 4/40 = 0.1

Since NaOH is completely dissociated, concentration of OH ions = 0.1 M = 10-1 mol L-1.

Kw=[H3O+][OH]

or 1 x 10-14 = [H3O+][OH]

or \(\frac{1 \times 10^{-14}}{10^{-1}}=\left[\mathrm{H}_3 \mathrm{O}^{+}\right]\)

or 1 x 10-13 M = [H3O+].

pH =- log[H3O+] = -log 10-13 = 13.

Example 14. Calculate the pH of a solution of a strong monobasic acid made by diluting 100 mL of a 0.01-M solution to 1 L
Solution:

M1V1 = M2V2

or 0.01 x100 = M2 x 1000.

∴ \(M_2=\frac{0.01 \times 100}{1000}=10^{-3}\)

Assuming complete dissociation

[H3O+] =10-3. (it is a monobasic add, one mole of protons is released per mole of add.)

∴ pH =- log[H3O+] =-[log10-3] = 3.

Example 15. Calculate the pH of a

  1. 0.02-M H2SO4 solution and
  2. 0.01-M Ba(OH)2 solution.

Solution:

1. The ionisation of H2SO4 can be represented by \(\mathrm{H}_2 \mathrm{SO}_4+2 \mathrm{H}_2 \mathrm{O} \longrightarrow 2 \mathrm{H}_3 \mathrm{O}^{+}+\mathrm{SO}_4^{2-}\)

Being a strong acid, H2SO4 is completely ionised.

1 mol of H2SO4 dissodates to give 2 mol of H3O+.

[H3O+] =2 x 0.02 = 0.04 mol L-1.

∴ pH = -log 0.04 =-log 4 x 10-2 = -(-1.3980) =1.398.

2. Ba(OH)2 dissodates completely.

1 mol of Ba(OH)2 dissociates to give 2 mol of OH.

∴ [OH] = 2[Ba(OH)2] =2 x 0.01 = 0.02 M.

Kw =[OH][H3O+]

∴ \(\left[\mathrm{H}_3 \mathrm{O}^{+}\right]=\frac{\mathrm{K}_{\mathrm{w}}}{\left[\mathrm{OH}^{-}\right]}=\frac{1 \times 10^{-14}}{0.02}=5 \times 10^{-13} \mathrm{M}\)

∴ pH =-log 5 x 10-13 =-(-12. 3010) =12.301

Example 16. How many grams of NaOH must be dissolved in 1 L of a solution of the base so that the pH of the solution is 13?
Solution:

pH = -log[H3O+]

[H3O+] = -antilog pH = -antilog13 =1 x 10-13 mol L-1.

Kw =[H3O+][OH].

∴ \(\left[\mathrm{OH}^{-}\right]=\frac{K_{\mathrm{w}}}{\left[\mathrm{H}_3 \mathrm{O}^{+}\right]}=\frac{1 \times 10^{-14}}{1 \times 10^{-13}}=1 \times 10^{-1} \mathrm{~mol} \mathrm{~L}^{-1}\).

As NaOH is a strong electrolyte, [NaOH] =[OH] =1 x 10-1 mol L-1.

Amount of NaOH = number of moles x mol. wt. of NaOH =1 x10-1 x 40 = 4 g.

Example 17. The dissociation constant of a weak acid HA is 1 x 10-9. Find the pH of a 0.1-M solution of the acid.
Solution:

The dissociation of the acid in water may be represented as follows.

⇒ \(\mathrm{HA}+\mathrm{H}_2 \mathrm{O} \rightleftharpoons \mathrm{H}_3 \mathrm{O}^{+}+\mathrm{A}^{-}\)

If x mol dissociates at equilibrium, [HA] = 0.1- x.

∴ \({\left[\mathrm{H}_3 \mathrm{O}^{+}\right]=\left[\mathrm{A}^{-1}\right]=x .}\)

∴ \(\mathrm{K}_{\mathrm{a}}=\frac{\left[\mathrm{H}_3 \mathrm{O}^{+}\right]\left[\mathrm{A}^{-}\right]}{[\mathrm{HA}]}=\frac{x \times x}{0.1-x}\).

Since x is negligible compared to 1,

∴ \(K_{\mathrm{a}}=\frac{x^2}{0.1}\)

∴ x = \(\sqrt{\mathrm{K}_{\mathrm{a}} \times 0.1}=\sqrt{1 \times 10^{-9} \times 0.1}=10^{-5} \mathrm{~mol} \mathrm{~L}^{-1}\)

∴ \({\left[\mathrm{H}_3 \mathrm{O}^{+}\right]=10^{-5} \mathrm{~mol} \mathrm{~L}^{-1} .}\)

∴ \(\mathrm{pH}=-\log \left[\mathrm{H}_3 \mathrm{O}^{+}\right]=-\log 10^{-5}=5\) .

Example 18. Calculate, the[H3O+] of a solution of a monobasic acid whose pH is 4.35.
Solution:

pH = -log[H3O+]

∴ log[H3O+] = -pH = -4.35.

[H3O+ ] = antilog(-4.35).

Since mantissas are always positive in logarithmic tables, add -1 to the characteristic and +1 to mantissa in order to make the mantissa positive.

[H3O+] = antilog[-4-1 + (-0.35 +1)] = antilog 5.65 = 4.467 x 10-5

Equilibrium  Multiple Choice Questions

Question 1. Which of the following is a characteristic of a reversible reaction?

  1. The number of moles of the reactants is equal to that of the products.
  2. It can be influenced by a catalyst.
  3. It can never proceed to completion.
  4. None of the above

Answer: 3. It can never proceed to completion.

Question 2. A chemical reaction A ⇔ B is said to be in equilibrium when

  1. A has been completely converted to b
  2. The conversion of a to b is 50%
  3. Only 10% of the conversion of a to b has taken place
  4. The rate of conversion of a to b is just equal to that of b to a

Answer: 4. The rate of conversion of a to b is just equal to that of b to a

Question 3. The equilibrium constants of these reactions are related as

⇒ latex]2 \mathrm{SO}_2(\mathrm{~g})+\mathrm{O}_2(\mathrm{~g}) \rightleftharpoons 2 \mathrm{SO}_3(\mathrm{~g}) K_1[/latex]

⇒ \(2 \mathrm{SO}_3(\mathrm{~g}) \rightleftharpoons 2 \mathrm{SO}_2(\mathrm{~g})+\mathrm{O}_2(\mathrm{~g}) K_2\)

  1. K2 = K1
  2. \(K_2=K_1^2\)
  3. \(K_2=\frac{1}{K_1^2}\)
  4. \(K_2=\frac{1}{K_1}\)

Answer: 4. \(K_2=\frac{1}{K_1}\)

Question 4. In which of the following cases does the reaction proceed the fastest?

  1. K =102
  2. K = 10-2
  3. K = 10
  4. K = 1

Answer: 1. K =102

Question 5. For the reaction \(\mathrm{PCl}_5(\mathrm{~g}) \rightleftharpoons \mathrm{PCl}_3(\mathrm{~g})+\mathrm{Cl}_2(\mathrm{~g})\), the forward reaction at constant temperature is favoured by

  1. The introduction of an inert gas at constant volume
  2. The introduction of an inert gas at constant pressure
  3. The introduction of Cl2 at constant volume
  4. An increase in pressure

Answer: 2. The introduction of an inert gas at constant pressure

Question 6. In the reaction \(2 \mathrm{~A}(\mathrm{~g})+\mathrm{B}(\mathrm{g}) \rightleftharpoons \mathrm{C}(\mathrm{g})+85 \mathrm{cal}\), which of the following conditions would give the best yield at equilibrium?

  1. 500 atm, 100°C
  2. 100 atm, 100°C
  3. 500 atm, 500°C
  4. 100 atm, 500°C

Answer: 1. 500 atm, 100°C

Question 7. Tire solubility of Agl in a solution of Nal is less than that in pure water because

  1. Agl forms a complex with nal
  2. Of the common-ion effect
  3. The Ksp of AgI is less than that of nal
  4. Of a decrease in temperature

Answer: 2. Of the common-ion effect

Question 8. Which of the following is not a Lewis acid?

  1. BF3
  2. AlCl3
  3. BeCl2
  4. SnCl4

Answer: 4. SnCl4

Question 9. An acid solution of pH 6 is diluted a hundred times. The pH of the solution becomes

  1. 6.95
  2. 6
  3. 4
  4. 9

Answer: 1. 6.95

Question 10. If the solubility of a salt M2X3 is x mol L-1, its solubility product is

  1. x5
  2. 76x2
  3. 96x5
  4. 108x2

Answer: 4. 108x2

Question 11. The solubility product of A2B is 4 x 10-9 mol L-1 Its solubility is

  1. 10-3 M
  2. 41/3 x 10-3 M
  3. 10-4 M
  4. 2 x 10-5 M

Answer: 1. 10-3 M

Question 12. Which of the following has the lowest value of Ksp at ordinary temperature?

  1. Mg(OH)2
  2. Ba(OH)2
  3. Ca(OH)2
  4. Be(OH)2

Answer: 4. Be(OH)2

Question 13. Which of the following is the strongest Bronsted base?

  1. ClO
  2. ClO2
  3. ClO3
  4. ClO4

Answer: 1. ClO

Question 14. The pH of a certain solution is 5.0. Just enough acid is added to decrease the pH to 2.0. The hydrogen ion concentration will

  1. Increase 1000 times
  2. Decrease 1000 times
  3. Increase 100 times
  4. Increase 10 times

Answer: 1. Increase 1000 times

Question 15. The Ksp of CaF2 is 1.7 x 10-10. The addition of what concentration of an equal volume of F ions to a solution containing 0.01-M Ca2+ ions will result in the precipitation of CaF2?

  1. 3.4×10-8 M
  2. 1.3×10-4 M
  3. 3.68 x 10-4 M
  4. 2.6×10-4 M

Answer: 2. 1.3×10-4 M

 

 

 

 

Thermodynamics – Definition, Equations, Laws

Thermodynamics

This chapter deals with energy changes that take place during a chemical reaction. In any chemical reaction, the atoms of the reactants are rearranged to form the products. This involves the breaking and forming of bonds. You know that energy is required to break bonds and that it is released when bonds are formed. It should then be easy to understand that a chemical reaction involving the dissociation and formation of bonds must be accompanied by energy changes.

The energy change accompanying a reaction may appear in different forms. When fuels are burnt, for instance, energy appears as heat and light. The chemical reaction in a battery produces electrical energy. When a grenade explodes, the chemical reaction produces heat, light, sound and kinetic energy. Consider the following reactions.

⇒ \(\mathrm{C}+\mathrm{O}_2\longrightarrow \mathrm{CO}_2 \text { + heat }\)

⇒ \(2 \mathrm{Mg}+\mathrm{O}_2\longrightarrow 2 \mathrm{MgO}+\text { heat }+ \text { light }\)

⇒ \(\mathrm{Zn}+\mathrm{CuSO}_4\longrightarrow \mathrm{ZnSO}_4+\mathrm{Cu}+\text { electrical energy }\)

In these reactions there is a net release of energy. There are other reactions in which energy is absorbed. Take the case of electrolysis, for example. Here electrical energy is absorbed and the electrolyte splits into its components. Photosynthesis, on which almost all living organisms depend directly or indirectly, is another example of a reaction in which energy is absorbed.

Thermodynamics Some Definitions

Before we start a systematic study of the energy changes associated with chemical reactions, let us define a few basic terms used in any discussion of energetics or chemical thermodynamics.

System: A system is that part of the universe which we are interested in investigating. For example, if we are studying a particular reaction in a vessel, then the vessel, the reactants, and the products constitute the system.

Surroundings: Everything other than the system or the part of the universe other than the system is called the surroundings. In the example we have just considered, everything other than the vessel in which the reaction is taking place is called the surroundings. A system is separated from the surroundings by boundaries which may be real or imaginary. Further, boundaries may be considered part of either the system or the surroundings, depending upon convenience.

There are three types of systems: open system, closed system, and isolated system.

Open system A system which can exchange matter and energy with the surroundings is called an open system. For example, an open test tube in which a reaction is taking place is an open system. It can exchange heat with the surroundings and gaseous products can escape into the surroundings.

Closed system A dosed system can exchange energy with the surroundings but not mass. If a reaction occurs in a sealed bulb, which can exchange heat with the surroundings but not matter, the bulb with the reactants and products constitutes a closed system.

Isolated system An isolated system can neither exchange matter nor energy with the surroundings. A thermos flask filled with hot tea is an isolated system.

Intensive and extensive properties: The properties of any substance can be classified as intensive or extensive. Intensive properties are those independent of the size of the substance. For example, by doubling the size of the given sample of a substance, the temperature and pressure of the substance do not double or do not change. These are called Intensive properties, other examples being viscosity, density, and all other molar properties.

If the value of the property depends on the size of the substance, it is called an extensive property. On doubling the size, internal energy doubles, and hence it is an extensive property. Other examples include mass, volume, heat capacity, enthalpy, entropy, and free energy.

State functions: State functions are those parameters or measurable macroscopic properties of a system that describe its state. The state of a system is defined by specifying the values of certain number of macroscopic properties. The number depends on the nature of the system.

  • The values of state functions or state variables depend only on the state of the system and not on how that state is readied. Take a gaseous system, for example. Its state is described by the state functions: temperature (T), pressure (p), and volume (IQ. To take a more specific case, consider 1 mol of CO at stp. Its volume will be 22.7 L irrespective of the method by which it is obtained.
  • All state functions or state variables are not independent because equations of state exist between different state functions. For example, consider the equation of state for an ideal gas, pV-nRT. Here out of the four variables (p. V. n and T), only three can be independently varied.
  • Thus once the values of the minimum number of state functions are fixed, others have definite values. The equilibrium state of a system is characterised by the definite values of state functions which do not change with time. Internal energy, enthalpy, and entropy are the state functions you will learn about in this chapter.

The state of a system: As stated above, thermodynamic equilibrium exists in a system when the macroscopic properties or state functions do not change with time. If the equilibrium state is disturbed tire system again settles down to a new equilibrium state.

  • If a particular state function does not have equal values in the system and its surroundings, exchange of matter or energy or both takes place between the thermodynamic system and its surroundings. After the interaction between system and its surroundings stops, the system attains a new equilibrium state with new values of state functions.
  • The starting state of the system is referred to as the initial state and the state reached after interaction with the surroundings is the final state. When a thermodynamic system undergoes a change of state (from the initial to tire final), we say it has undergone a process. This implies that there will be a change in at least one of tire state functions of tire system during a process.
  • A system can change from one state (initial state) to another (final state) through a number of paths. In other words, there can be a number of processes between an initial state and a final state. There are certain processes in which a particular state variable remains unchanged. Let us learn about these processes.

Isothermal If the temperature of the system remains constant during the change, the process is called isothermal.

Adiabatic If the system does not exchange heat with the surroundings, the process is called adiabatic.

Isobaric If the pressure of the system remains constant during the change, the process is called isobaric.

Isochoric If the volume of the system remains constant during the change, the process is called isochoric.

Reversible process In this process, the initial and the final states are connected through a succession of equilibrium states, i.e., at every state along the reversible path, there exists an equilibrium between the system and the surroundings. Such states are called quasi-equilibrium states.

Irreversible process The processes occurring in nature are generally irreversible. Their only difference from a reversible process is that equilibrium is not maintained during the transformation process.

Cyclic process The process that brings back a system to its original state after a series of changes is called a cyclic process.

Transference Of Energy

While describing open, closed, and isolated systems, we touched upon the idea of a system exchanging energy with the surroundings. There are two forms in which energy is exchanged between a system and the surroundings.

We often talk about heat flowing from a body at a higher temperature to one at a lower temperature. What we mean is that energy is exchanged between the two bodies because of a difference in temperature.

  • So, one of the forms in which a system can exchange energy with the surroundings is heat. Energy is exchanged between a system and the surroundings in the form of heat when they are at different temperatures.
  • The other mode of energy exchange between a system and the surroundings is energy. One example where energy is exchanged in this form is when the system and surroundings are at different pressures.
  • Consider a gas enclosed in a cylinder that is fitted with a weightless, frictionless moving piston. If the gaseous system is at a higher pressure than the surroundings, the piston will move upwards (increasing the volume of the gas) until the pressure of the system and that of the surroundings become equal.

During the expansion of the gas, work is done by the system on the surroundings due to the pressure difference between the two. Here energy is transferred from the system to the surroundings as work.

Basic Chemistry Class 11 Chapter 6 Thermodynamics When The Gas Enclosed In A Cylinder Fitted With A Piston Is AT A Higher Pressure Than The Surroundings

  • If, on the other hand, the pressure of the system is lower than that of the surroundings, the piston is pushed down until the pressures become the same. In this case the volume of the gas decreases and work is done by the surroundings on the system.
  • Once again there is figure, when the gas enclosed in a cylinder transference of energy in the form of work, but from the fitted with a piston is at a higher pressure than the surroundings to the system. By convention, work done on the surroundings, the piston moves until the pressures system is positive, while work done by the system is negative. equalize.

Remember that neither heat, nor work is a property of a system. They are path functions as they depend on how a change is brought about. A system does not possess a particular amount of work or heat. It does, however, have a particular amount of energy under a particular set of conditions. The SI unit of heat is the same as that of work, i.e., the joule. However, heat is sometimes expressed in calories or kilocalories.

1 calorie = 4.184 J.

1 J = 107 ergs.

Work

There are several kinds of work, such as mechanical work, electrical work, and chemical work. Work is said to be done if an object is moved by a force F through a distance d. The work done is given by

W = Fxd…….Equation 1

Some types of work encountered in chemistry are:

  1. Expansion work (which we shall study in detail) in the case of expansion or compression of a gas against a force surface expansion in case of liquids—these are examples of mechanical work
  2. When larger molecules are synthesized from smaller ones within living organisms—chemical work
  3. An ion moving in the presence of an electric field—electrical work.

The most common form of mechanical work encountered in chemistry is the expansion work associated with the expansion of a gas against pressure. It is also called p-V or pressure-volume work.

Basic Chemistry Class 11 Chapter 6 Thermodynamics Work Is Done On The System

Basic Chemistry Class 11 Chapter 6 Thermodynamics Work Is Done By The System And If The Weight Is Lowered

To understand the concept further, let us once again consider a Figure but with a weight attached to the piston. When the gas expands, the piston moves up and the weight is lifted up. We say that work is done by the system.

Let us now calculate the work done in terms of pressure and volume. If an external pressure pex is applied on the piston and the piston moves inwards by a distance dl, then the change in volume of the gas is given by

dV=A • dl …….(1)

where A is the area of the cross-section of the piston. The force on the piston is given as

F = pex • A …..(2)

Now, the work done in moving the piston by a distance dl against an opposing force of magnitude F is given by

dW = -Fdl …….(3)

(the negative sign is due to the opposing nature of the force).

Substituting (2)  in (3)

dW = -pex • Adl

or dW = -pex dV

or simply W = -p Δ V …..Equation 2

  • By the help of this equation, we can calculate the work done (or pressure/volume) during the expansion or compression of a gas. When expansion occurs ΔV in Equation 2 is positive, thereby resulting in a negative value for W.
  • Similarly, when compression occurs, weight is lowered in the surroundings and we say that work is done on the system. Since AV is negative now, W is positive.
  • Thus we can say that work is done when there is a change in height of a weight in the surroundings.

If due to a change in state of a system, a weight is lifted in the surroundings, work is done by the system and if a weight is lowered in the surroundings, work is done on the system. Now W can also be given as W = mgh,

where m is the mass lifted, g is the acceleration due to gravity and h is the height through which the mass is lifted.

Thus, during a change in state of the system, we can learn about any work done by measuring the relevant quantities only in the surroundings.

Reversible expansion or compression of a gas: Suppose a gas confined in a cylinder with a movable piston is in equilibrium. This means that there is no change in state of the system. This is possible when the external pressure pex (pressure that is applied) is equal to the internal pressure, p, of the gas. A small change in pex can disturb the equilibrium and the gas may expand or compress, resulting in a change in the volume and hence the state of the system.

  • When pex is increased infinitesimally (infinitesimal means negligibly small) then compression occurs and the volume of the gas decreases. Equilibrium with the surroundings is also established almost immediately. The system can also be restored to its original state by reducing the external pressure to the original value.
  • By continuously increasing the external pressure, infinitesimally, the gas can be made to undergo a finite amount of compression. This is achieved by keeping different weights on the piston one after the other.
  • In each step, however, the change is infinitesimal and can be reversed by an infinitesimal change in external pressure. Again, in each step equilibrium is immediately established.
  • Thus we have a process which is carried out in several infinitesimally small steps and throughout the process the system is in equilibrium with its surroundings. This is a reversible process.

If the difference between the internal pressure and the external pressure is large then there is a considerable disturbance of equilibrium and an infinitesimal change in pex in the opposite direction will not bring back the system to its original state. Such a process is termed an irreversible process.

Calculation of work done in various cases

1. Work done in free expansion When the external pressure is zero (pex = 0), there is no opposing force on the piston and the gas can expand freely. This is called free expansion. This occurs when a gas expands in vacuum. Both the work done on the system and by the system are zero in this case.

2. Work done in expansion against constant pressure If a constant external pressure is applied, the gas expands or compresses until the internal pressure becomes equal to the external pressure. The work done at every displacement dV is given by Equation 2. The total work done in the expansion from V1 to V2 is the sum of all such contributions. It is obtained by finding the area under the p-V plot, i.e., by integrating Equation.

∴ \(W=-\int_{V_1}^{V_2} p_{e x} d V\) ….. (1)

= \(-p_{e x} \int_{V_1}^{V_2} d V=-p_{e x}\left(V_2-V_1\right)\)

∴ \(W=-p_{e x} \Delta V\)

In case of expansion, ΔV is positive and W turns out to be negative. Therefore, work is done by the system.

If ΔV is negative as in case of compression, W is positive, i.e., work is done on the system.

Basic Chemistry Class 11 Chapter 6 Thermodynamics Plot Of p Versus V For Expansion Against Constant Pressure

3. Work done in reversible expansion In reversible expansion, the difference between external and internal pressure is infinitesimally small. With very small changes in pressure, the process of expansion and compression may be reversed. If the expansion occurs in several stages, at each stage, it is ensured that the external pressure is only infinitesimally different (less in case of expansion and more in case of compression) from the internal pressure of the gas.

Since \(p_{e x}=p_{\text {in }}\) we may replace pex in Equation 2 by pin

dW = -pin dV.

This is the work done at every stage of the expansion. The total work is obtained by summing up all such infinitesimal contributions (again) given by the area under the p-V plot.

⇒ \(\int d W=-\int p_{\mathrm{ma}} d V\)

∴ \(W_{\mathrm{rev}}=-\int p_{\mathrm{in}} d V\)

The subscript ‘rev’ indicates the reversible nature of the process.

If it is assumed that the gas behaves ideally then pin be replaced by nRT/v using the ideal gas equation.

Therefore \(W_{\mathrm{rev}}=-\int_{V_1}^{V_2} \frac{n R T}{V} d V\)

If the process is carried out at a constant temperature (isothermally), T may be taken out of the integral sign.

⇒ \(W_{r e v}=-n R T \int_{V_1}^{V_1} \frac{d V}{V}\)

= \(-n R T|\ln V|_{V_1}^{V_2}\)

or, \(W_{n e v}=-n R T \ln \left(\frac{V_2}{V_1}\right)=-2303 n R T \log \frac{V_2}{V_1}\),

where V1 and V2 are the initial and final volumes respectively.

Basic Chemistry Class 11 Chapter 6 Thermodynamics Plot Of P Versus V For A Reversible Expansion

Internal Energy

The energy stored within a system (say chemical system) is its internal energy or intrinsic energy. Under a particular set of conditions, a thermodynamic system has a definite amount of internal energy.

  • In other words, the internal energy of a system, generally represented as U, depends on the nature, amount, temperature, and pressure of the system. This energy possessed by a system is due to the different types of energy that its atoms or molecules have.
  • It is, in fact, the sum of the energies of the elementary particles of the system, i.e., the sum of their potential and kinetic energies and the bond energy between constituting atoms.
  • Instead of saying that a particular system under a particular set of conditions possesses a definite amount of internal energy, we could have said that internal energy is a state function, which means that it depends only on the initial and final state of the system and is independent of the way (path) the change takes place.

However, unlike other state functions, like pressure and temperature, it is not possible to determine the value of internal energy. This is because it is not possible to determine with exactness the values of the components of internal energy. Note that this matters little, because in the study of chemical processes, what we are really interested in knowing is the clumge in internal energy, ΔU (delta U).

ΔU = U2-U1

where U2 is the internal energy of the system at the end of the process being studied and is the internal energy of the system at the start of the process. For a chemical reaction

ΔU = Up-Ur,

where U2 is the internal energy of the products and Ur is that of the reactants. Obviously, All is a state function since Up and Ur are state functions.

So, how do we measure ΔU or the change in internal energy associated with a chemical process? Suppose we carry out a chemical reaction in such a way that the system remains at the same temperature, there is no work done on the system and it does no work on the surroundings.

  • Whatever the change in the internal energy of the system in the course of the reaction must be equal to the energy exchanged by the system with the surroundings in the form of heat.
  • Why? This in fact follows from the law of conservation of energy, which we will discuss subsequently, for the moment, imagine that the change in internal energy of the system is negative, i.e., Up < Ur.
  • What happens to this energy which the system has apparently lost? It cannot just vanish because that is against the law of conservation of energy. We have assumed that the system does not work and nor is any work done on it (i.e., volume remains constant), so the energy cannot be exchanged with the surroundings in the form of work.
  • Since the temperature is constant, the energy must be exchanged with the surroundings in the form of heat.
  • The change in the internal energy associated with a chemical reaction is, thus, determined by letting the reaction occur at constant temperature and constant volume and measuring the heat exchanged with the surroundings.

Law Of Conservation Of Energy

This is one of those common-sense laws which seem very obvious but have far-reaching implications. It states that energy can neither be created nor destroyed, though it may change from one form to another.

In the context of the system and the surroundings that we have been discussing, this means that the total energy of the system and the surroundings (i.e., the universe) remains constant.

  • More specifically, it means that during a chemical reaction, energy may be absorbed or released, but the total energy of the reaction system and the surroundings remains constant.
  • The law of conservation of energy is also known as the first law of thermodynamics. It is of relevance not only to chemists but also to physicists, engineers, and others who deal with the conversion of energy from one form into another.

The first law of thermodynamics being identical to the law of conservation of energy can be stated as:

  1. Energy can neither be created nor destroyed although it can be changed from one form to another.
  2. The energy of an isolated system is constant.

Consider the universe, which is an isolated system. Tire energy of the universe is conserved. Inside the universe, energy can be transferred from one part to another or it can be converted from one form to another, but it can neither be created nor destroyed.

Let us see if we can express this law mathematically. If the internal energy of a particular system is U1 and it absorbs a certain amount of heat q, then its internal energy will become U1 + q. Now suppose W amount of work is done on the system, then U2, the final energy of the system is

U2 =U1 +q + W

or ΔU = U2 – U1 = q + W….. Equation 3

(If W is the work done on the system, -W is the work done by the system.)

  • This means the change in the internal energy of the system is equal to the sum of the heat absorbed by the system and the work done on it.
  • One can generalise this to say the change in the internal energy of the system is the sum of the heat exchanged by the system (with the surroundings) and the work done on or by the system. Equation 3 is the mathematical statement of the first law of thermodynamics.

If we now substitute for W in Equation 3, the expression for change in internal energy (when only pressure-volume type of work is done) becomes

ΔU = q – pΔV …. Equation 4

Now, if there is no change in the volume of the system then W = 0, i.e., W = -∫pdV = 0.

We can easily see that if there is no change in volume during a reaction, Equation 4 becomes ΔU = qv,

where qv is the heat exchanged (evolved or absorbed) by the reaction system at constant volume. When heat is absorbed by the system, q is positive and when heat is evolved by the system, q is negative. This is exactly what we said in the previous section.

The change in the internal energy of a system during a chemical reaction is equal to the heat exchanged with the surroundings, provided the volume and temperature of the system remain constant.

For any isothermal expansion of an ideal gas the total energy remains the same, and q = -W. In other words, the internal energy remains constant (ΔU = 0) when an ideal gas expands isothermally.

Thus for an isothermal irreversible change, q = -W = pex ΔV

and for an isothermal reversible change, q = \(-W=n R T \ln \frac{V_2}{V_1}\)

= \(-2.303 n R T \log \frac{V_2}{V_1}\),

where V1 and V2 are the initial and final volumes respectively.

For an adiabatic change, q = 0,

∴ ΔU = Wadiabstic  (from Equation 3).

Enthalpy

So far we have considered reactions taking place at constant volume and temperature, and the energy changes associated with such reactions. But the reactions we carry out in the laboratory normally occur in open vessels (beakers or test tubes, say).

Obviously, the volume of such a reaction system does not remain constant. However, the pressure does, since such a system, open to the atmosphere, is at atmospheric pressure and this pressure is more or less constant.

  • The heat exchanged with the surroundings by a reaction system in the course of a reaction is equal to the change in the internal energy of the system if the temperature and volume are constant.
  • What about the heat exchanged with the surroundings by a reaction system at constant pressure and not maintained at constant volume? It should be different, but why and how?
  • Let us consider a reaction in which the volume of the system increases. An increase in volume can occur when the system does some work against the atmospheric pressure and energy is required for this work.
  • Thus, in this case, the heat exchanged with the surroundings would be lower than that for a system at constant volume and temperature.
  • On the other hand, if the reaction involves a decrease in the volume of the system, work would have to be done on the system by tire surroundings. The heat exchanged would then be greater than that exchanged by a system at constant volume.
  • Thus, when a reaction proceeds at constant pressure and temperature, the heat exchanged by the system with tire surroundings is not equal to the change in the internal energy of the system.

In other words, the change in the internal energy of the system is not the only contributing factor to the total energy change associated with the reaction. Tire total energy change includes change in energy due to the pressure-volume type of work done by or on tire system.

Enthalpy or heat content is a property or state function introduced to take care of the energy changes associated with tire kind of system we have just been discussing. It is denoted by H and expressed mathematically as follows.

H = U + pV, …… Equation 5

where U is the internal energy, p the pressure, and V the volume of the system. The enthalpy of a system (substance) can be defined as the total energy associated with it, i.e., its internal energy and the energy due to factors such as pressure-volume conditions. The enthalpy of a substance depends on its state (temperature, pressure, etc.).

  • Enthalpy changes are normally expressed with the substances in their standard states. The standard state of a pure substance is the pure form at a pressure of 1 bar and a specified temperature, the conventional temperature being 298.15 K.
  • Enthalpy change in the standard state is called the standard enthalpy change. It is denoted by \(\Delta H^ominus\), where the superscript \(\ominus\) denotes standard conditions.

The molar enthalpy of a substance Hm = H/n (where n is the number of moles) is an intensive property and Hm = Um + pVm. When enthalpy changes have to be compared, the states of the systems must be identical.

Enthalpy change: As in the case of internal energy, the absolute value of the enthalpy of a system cannot be determined. What we are interested in knowing is the enthalpy change associated with a process. The total change in the energy of a system during a reaction at constant pressure is the enthalpy change, denoted by ΔH.

∴ \(\Delta H=H_{\text {products }}-H_{\text {reactants }}\)

= \(\left(U_p+p V_p\right)-\left(U_r+p V_r\right)\)

= \(\Delta U+\Delta p V\),

where Up = internal energy of products,

Vp =volume of products,

Ur = internal energy of reactants, and Vt = volume of reactants.

Since p is constant, ΔH = ΔU + pΔV…….. Equation 6

The enthalpy change during a process is the sum of the change in internal energy and the pressure-volume work done.

Suppose the heat exchanged by a system with the surroundings during a reaction at constant pressure and temperature is qp. Suppose also that the change in internal energy of the system is ΔU and the work done on the system is W. Then

qp = ΔU-W = ΔU – (-pΔV)

= ΔU + pΔV

= ΔH. (ΔU + pΔV = ΔH)

  • The heat exchanged by a system with the surroundings during a reaction at constant pressure and temperature is the enthalpy change associated with the reaction. In practice, the enthalpy change associated with a reaction is determined by insulating the system and allowing the heat of the reaction to alter its temperature.
  • The amount of heat required to be supplied to or taken away from the system to let its temperature go back to the original value is then calculated to obtain the enthalpy change.

Significance of ΔU and ΔH: It is understood that the amount of heat exchanged with the surroundings for a reaction at a constant pressure (ΔH) is different from that exchanged at constant volume and temperature (ΔU).

The difference is not significant for solid or liquid systems but it does matter when gases are involved. Let us consider a gaseous reaction, where Vr is the total volume of the gaseous reactants Vp is the total volume of gaseous products, nr, is the number of moles of reactants and nr is the number of moles of products, all at constant pressure and temperature.

By the ideal gas law,

∴ \(\quad p V_t=n_r R T\)

and \(p V_p=n_p R T\)

Thus, \(p V_p-p V_r=\left(n_p-n_r\right) R T\)

or \(\quad p\left(V_p-V_r\right)=\left(n_p-n_r\right) R T\)

or, \(\quad p \Delta V=\Delta n_g R T\).

Here, Δng is the difference between the number of moles of the gaseous products and that of the gaseous reactants.

From the definition of enthalpy, it follows that ΔH = ΔU + pΔV.

Therefore, ΔH = ΔU + ΔngRT.

The above equation is useful in situations where one of the two quantities ΔH and ΔU is known and the other has to be calculated. ΔH and ΔU differ significantly for processes involving gases. For solids and liquids, pVp is only slightly different from pVx.

Source of enthalpy change: You may be wondering what exactly leads to enthalpy change during a reaction. Any reaction involves the breaking and forming of bonds.

  • Energy is released when bonds are broken and required for the formation of bonds. Simply put, the net change in energy due to the breaking and forming of bonds is the enthalpy change of the reaction.
  • Of course, for the net change in energy to be termed enthalpy change, the reaction must take place at constant pressure and temperature. The simplest case is that of a gaseous system in which gas A reacts with gas B to form gas C.

If the reaction involves solutions, the interactions (hence energy changes) between the solvents and the reactants and products have to be considered. For liquid or solid reactants, we have to also consider the interactions between the molecules.

Enthalpy change(for a gaseous system at constant pressure) = (energy required to break bonds) – (energy released in formation of bonds).

Let us consider the reaction between hydrogen gas and chlorine gas to form HCl gas. Energy is required to break tire H—H and Cl—Cl bonds and released when H—Cl bonds are formed.

∴ \(\mathrm{H}_2(\mathrm{~g})+\mathrm{Cl}_2(\mathrm{~g}) \longrightarrow 2 \mathrm{HCl}(\mathrm{g})\)

The energy required to break H—H bonds = 436 kJ mol-1.

Energy required to break Cl—Cl bonds = 242 kJ mol-1.

Energy released in the formation of two H—Cl bonds = 2 x 431 kj mol-1.

Enthalpy change (ΔH) = 436 + 242 – 2 x 431 = -184 kJ.

The negative value of ΔH indicates a decrease in the enthalpy of the system.

The heat evolved in this reaction is 184 kJ mol-1.

Exothermic And Endothermic Reactions

In our discussion of the energy changes associated with chemical reactions so far, we have only mentioned that energy is released in some reactions and absorbed in others. Reactions in which energy is released are called exothermic while those in which energy is absorbed are called endothermic.

Exothermic reactions: While writing an exothermic reaction, the heat evolved is indicated on the right side, after the products.

∴ \(\mathrm{N}_2+3 \mathrm{H}_2 \longrightarrow 2 \mathrm{NH}_3+93.7 \mathrm{~kJ}\)

  • If an exothermic reaction is carried out at constant volume and temperature, the heat evolved (qν) is (numerically) equal to the change in internal energy (ΔU). In such a reaction the internal energy of the products (Up) is less than the internal energy of the reactants (Ur) and ΔU is negative.
  • If an exothermic reaction occurs at constant pressure and temperature, the heat evolved is (numerically) equal to the change in enthalpy (ΔH). The enthalpy of the products is less than the enthalpy of the reactants, i.e., ΔH is negative.

Basic Chemistry Class 11 Chapter 6 Thermodynamics Graphical Representation Of Enthalpy Change In An Endothermic Reaction

Basic Chemistry Class 11 Chapter 6 Thermodynamics Graphical Representation Of Enthalpy Change In An Exothermic Reaction

Endothermic reactions: In the case of an endothermic reaction, the heat absorbed can be indicated along with reactants or with the products. Obviously, if it is shown on the left side of the equation, it will have a positive sign, and if it is shown on the right side, it will carry a negative sign.

∴ \(\mathrm{N}_2+\mathrm{O}_2+180.5 \mathrm{~kJ} \longrightarrow 2 \mathrm{NO}\)

or, \(\mathrm{N}_2+\mathrm{O}_2 \longrightarrow 2 \mathrm{NO}-180.5 \mathrm{~kJ}\)

When an endothermic reaction occurs at constant temperature and constant volume, the heat absorbed is (numerically) equal to the change in the internal energy of the system. The internal energy of the products (Up) is greater than the internal energy of the reactants (Ur) and ΔU is positive.

When an endothermic reaction proceeds at constant pressure and temperature, the heat absorbed is (numerically) equal to the change in enthalpy of the system. The enthalpy of the products (Hp) is greater than the enthalpy of the reactants (Hr) and ΔH is positive.

Thermochemical Equations

You already know how to write a chemical equation. When a chemical equation not only indicates the quantities and physical states of the reactants and products involved, but also the change in enthalpy during a reaction, it is called a thermochemical equation. Fractional coefficients may also be used in such an equation.

⇒ \(2 \mathrm{H}_2(\mathrm{~g})+\mathrm{O}_2(\mathrm{~g}) \longrightarrow 2 \mathrm{H}_2 \mathrm{O}(\mathrm{l}) ; \quad \Delta H=572 \mathrm{~kJ}\)

Certain conventions must be followed while writing a thermochemical equation. These conventions are listed below.

1. The heat evolved or absorbed is indicated as in the equation above, in terms of ΔH. Remember that ΔH is negative for exothermic reactions and positive for endothermic reactions.

2. The numerical value of ΔH corresponds to the reaction as written. In the absence of any information, it is assumed that a certain value of ΔH is due to the number of moles of reactants that combine as indicated by the chemical equation. Thus ΔH is expressed in kJ mol-1.

3. The value of standard enthalpy (\(\Delta H^{\ominus}\)) in a thermochemical equation corresponds to the standard state of the substances involved in the reaction. The term ‘standard state of a substance’ refers to the pure substance at exactly 1 bar pressure.

4. The coefficients of the various substances represent the number of moles of the substances involved in the reaction and the value of ΔH corresponds to these coefficients. It stands to reason, therefore, that if the coefficients are multiplied or divided by some factor, so must ΔH be. Consider the following example.

⇒ \(\mathrm{H}_2(\mathrm{~g})+\frac{1}{2} \mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{H}_2 \mathrm{O}(\mathrm{g}) \quad \Delta H=-242 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

If the coefficients are multiplied by 2, ΔH must also be multiplied by 2.

⇒ \(2 \mathrm{H}_2(\mathrm{~g})+\mathrm{O}_2(\mathrm{~g}) \longrightarrow 2 \mathrm{H}_2 \mathrm{O}(\mathrm{g}) \quad \Delta \mathrm{H}=-(242 \times 2) \mathrm{kJ} \mathrm{mol}^{-1}\)

When the reaction is reversed, the sign of AH is reversed but its magnitude remains the same.

\(\mathrm{H}_2(\mathrm{~g})+\mathrm{I}_2(\mathrm{~g}) \longrightarrow 2 \mathrm{HI}(\mathrm{g})\Delta H=+53.9 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

⇒ \(2 \mathrm{HI}(\mathrm{g}) \longrightarrow \mathrm{H}_2(\mathrm{~g})+\mathrm{I}_2(\mathrm{~g}) \Delta H=-53.9 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

5. The physical states of the reactants and products have to be indicated because ΔH changes with the physical state. The following example will make this clear.

⇒ \(\mathrm{H}_2(\mathrm{~g})+\frac{1}{2} \mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{H}_2 \mathrm{O}(\mathrm{g})\Delta H=-242 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

⇒ \(\mathrm{H}_2(\mathrm{~g})+\frac{1}{2} \mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{H}_2 \mathrm{O}(\mathrm{l})\Delta H=-286 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

Enthalpy Of Reactions

The enthalpy change accompanying a reaction is called reaction enthalpy and is denoted by ΔrH. The standard enthalpy of reaction (ΔrHΘ) is the change in enthalpy per unit amount of the reaction when the reactants in their standard states change to products in their standard states.

By standard states we mean reference state, i.e., the most stable state of aggregation. For example the reference state of H2 is pure gas at 1 bar and that of CaCO3 is pure solid at 1 bar. By convention the standard states are reported at 298 K. There are several factors that determine the value of enthalpy of a reaction.

Quantities of reactants The amount of heat absorbed or released as also the enthalpy of a reaction depends on the quantities of reactants involved. This should be pretty easy to understand. If you bum 1 kg of coal it will produce less heat than if you bum 10 kg of coal.

The standard enthalpy change for the reaction \(\mathrm{C}(\mathrm{s})+\mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{CO}_2(\mathrm{~g}) \quad \Delta_{\mathrm{r}} \mathrm{H}^{\ominus}=-393.5 \mathrm{~kJ} \mathrm{~mol}^{-1}\) has the units of kJ mol-1. Per mole refers to the reaction as written above.

Let us take another example.

⇒ \(2 \mathrm{CH}_3 \mathrm{OH}(\mathrm{l})+3 \mathrm{O}_2(\mathrm{~g}) \longrightarrow 2 \mathrm{CO}_2(\mathrm{~g})+4 \mathrm{H}_2 \mathrm{O}(\mathrm{g})\)

∴ \(\Delta_{\mathrm{r}} H^{\ominus}\) for this reaction is -1276.9 kj mol-1.

However, if the reaction is written with different stoichiometric coefficients, i.e., \(\mathrm{CH}_3 \mathrm{OH}(\mathrm{l})+\frac{3}{2} \mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{CO}_2(\mathrm{~g})+2 \mathrm{H}_2 \mathrm{O}(\mathrm{g})\)

∴ \(\Delta_{\mathrm{r}} H^{\ominus}\) is -638.4 kJ mol-1, ‘per mole’ referring to the reaction as written here.

Physical State of Reactants and products The change of state of a substance involves heat changes, so the enthalpy of a reaction depends on the physical states of the reactants and products. We have already discussed the case of the formation of liquid water and gaseous water vapour from gaseous hydrogen and oxygen.

Allotropic forms The enthalpy of a reaction also depends on the allotropic forms of the substances involved in the reaction.

⇒ \(\mathrm{C} \text { (diamond) }+\mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{CO}_2(\mathrm{~g}) \Delta H^{\ominus}=-395.4 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

⇒ \(\mathrm{C} \text { (graphite) }+\mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{CO}_2(\mathrm{~g}) \Delta H^{\ominus}=-393.5 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

Temperature The enthalpy of a reaction depends on the temperature at which the reaction proceeds. The value of enthalpy indicated in a thermochemical equation is generally the value at 298 K.

Constant pressure or volume As we have discussed already, the enthalpy of a reaction at a particular temperature depends on whether the reaction occurs at constant pressure or constant volume. From our discussion so far

ΔH = ΔU + pΔV.

But ΔH = qp and ΔU = qv.

∴ qp = qv + pΔV

Hess’s Law

The enthalpy of a substance is a stale function, independent of the method by which the substance is made. The enthalpy change in a reaction is the difference between the enthalpies of the products and the reactants.

  • Both of these are consequences of the law of conservation of energy. Another consequence of the law is what is known as
    Hess’s law of constant heat summation, which states that the enthalpy change in a chemical or physical process is independent of the path taken or the manner in which the change is brought about.
  • In the context of a chemical reaction, the law can be stated as follows. “The amount of heat absorbed or released during a reaction is the same whether the reaction proceeds in a single step or through several steps.”

Though G H Hess (a Russian chemist) came up with this law as a result of experimental observations, if you think a little you will realise that it follows from, or is a corollary of, the law of conservation of energy. How? Suppose a reactant A is converted into a product D and the heat evolved in the process is q.

  • Now suppose the same reactant A changes first to B, then B is converted to C and finally, C changes to D and the heat evolved in the three steps is q1, q2, and q3. Let q1 + q2 + q3 = q′.  According to Hess’s law q = q’. If Hess’s law were not true, either q> q’ or q < q’. Let us consider the possibility that q >q’.
  • Then the heat evolved in converting A to D in a single step would be greater than the heat absorbed when D is converted to A in three steps. This would lead to a situation in which (q – q‘) of heat would be ‘created’ after the completion of the cyclic process and the law of conservation of energy would be violated.

Basic Chemistry Class 11 Chapter 6 Thermodynamics A Is Converted Into D In A Single Step Or Through Three Steps

Hess’s law can be stated as follows: The enthalpy change for a reaction that is the sum of two or more other reactions is equal to the sum of the enthalpy changes of the constituent reactions.

Let us consider two ways in which carbon dioxide may be produced from carbon and oxygen. Either carbon dioxide can be produced directly by the combustion of carbon, or carbon can first be converted to carbon monoxide, which can then be converted to carbon dioxide.

Basic Chemistry Class 11 Chapter 6 Thermodynamics Hess's Law

⇒ \(\mathrm{C}(\mathrm{s})+\mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{CO}_2(\mathrm{~g}) \Delta_{\mathrm{r}} H^{\ominus}=-394 \mathrm{~kJ}\) …. (1)

⇒ \(\mathrm{C}(\mathrm{s})+\frac{1}{2} \mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{CO}(\mathrm{g}) \Delta_{\mathrm{r}} H_1^{\ominus}=-110.5 \mathrm{~kJ}\) …. (2)

⇒ \(\mathrm{CO}(\mathrm{g})+\frac{1}{2} \mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{CO}_2(\mathrm{~g}) \Delta_{\mathrm{r}} H_2^{\ominus}=-283.5 \mathrm{~kJ}\) …..(3)

Adding the enthalpy changes in (2) and (3) you will get -394 kJ mol-1, which is the same as the enthalpy change in (1). Do you see now why the law is called Hess’s law of constant heat summation?

Application of Hess’s law: Hess’s law is very useful in determining enthalpy changes in reactions for which it is not possible to experimentally determine enthalpy changes. It follows from Hess’s law that thermochemical equations can be added, subtracted, multiplied, or divided like algebraic equations.

Let us again consider the reaction between carbon and oxygen. The principal product of this reaction is CO2 but CO may also be produced due to insufficient oxygen and this may then form CO2. It is difficult to measure directly the enthalpy of combustion of carbon to carbon monoxide, CO. Let us assume it is unknown and determine its values by applying Hess’s law.

Reaction for which enthalpy is to be found: \(2 \mathrm{C}(\mathrm{s})+\mathrm{O}_2(\mathrm{~g}) \longrightarrow 2 \mathrm{CO}(\mathrm{g})\) …(1)

Reactions for which data is available: \(\mathrm{C}(\mathrm{s})+\mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{CO}_2(\mathrm{~g})\) \(\Delta_{\mathrm{r}} H^\ominus=-393.5 \mathrm{~kJ}\) …. (2)

⇒ \(2 \mathrm{CO}(\mathrm{g})+\mathrm{O}_2(\mathrm{~g}) \longrightarrow 2 \mathrm{CO}_2(\mathrm{~g})\) \(\Delta_{\mathrm{r}} H^{\ominus}=-566.0 \mathrm{~kJ}\) ….. (3)

How do we use these equations (with data) to get equation (1)? We need C and O2, on the LHS and CO on the RHS. In Equation (2) C and O2 are already on the LHS but the coefficient of C is 1 while we require 2 (from Equation (1)).

Hence, by multiplying Equation (2) by 2, we get

⇒ \(2 \mathrm{C}(\mathrm{s})+2 \mathrm{O}_2(\mathrm{~g}) \longrightarrow 2 \mathrm{CO}_2(\mathrm{~g}) \quad \Delta_{\mathrm{r}} H^{\ominus}=2 \times(-393.5) \mathrm{kJ}\) ….(4)

[If the equation is multiplied by 2, \(\Delta_r H^{\ominus}\)must also be multiplied by 2.]

Next CO must appear on the RHS. Reversing Equation (3) to get CO on RHS, we get

⇒ \(2 \mathrm{CO}_2(\mathrm{~g}) \longrightarrow 2 \mathrm{CO}(\mathrm{g})+\mathrm{O}_2(\mathrm{~g}) \quad \Delta_{\mathrm{r}} H^{\ominus}=-(-566)=+566 \mathrm{~kJ}\) …. (5)

[If the reaction is reversed, the sign of ΔH must also be reversed.]

Now adding reactions (4) and (5)

∴ \(2 \mathrm{C}(\mathrm{s})+2 \mathrm{O}_2(\mathrm{~g}) \longrightarrow 2 \mathrm{CO}_2(\mathrm{~g}) \quad \Delta_{\mathrm{r}} H^{\ominus}=-787.0 \mathrm{~kJ} \\
\begin{array}{lll}
2 \mathrm{CO}_2(\mathrm{~g}) & \longrightarrow 2 \mathrm{CO}(\mathrm{g})+\mathrm{O}_2(\mathrm{~g})  \Delta_{\mathrm{r}} H^{\ominus}=566 \mathrm{~kJ} \\
\hline 2 \mathrm{C}+\mathrm{O}_2 \longrightarrow 2 \mathrm{CO} & \Delta H^{\ominus}=-787+566=-221 \mathrm{~kJ}
\end{array}\)

Basic Chemistry Class 11 Chapter 6 Thermodynamics Calculation Of Enthalpy Of partial Combustion Of Carbon To Give Carbon Monoxide

Enthalpy Of Formation

The enthalpy or heat of the formation of a substance is the heat change accompanying the formation of 1 mol of the substance from its constituent elements and is usually written as ΔtH. If all the substances involved in the reaction are in the standard state, the heat change is referred to as the standard heat of formation and is denoted by \(\quad \Delta_{\mathrm{f}} H^{\ominus}\) Consider the following equations.

∴ \(\mathrm{C}(\mathrm{s})+\mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{CO}_2(\mathrm{~g}) \quad \Delta_{\mathrm{f}} H^{\ominus}=-393.5 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

This means the heat liberated in the formation of 1 mol of gaseous carbon dioxide from its constituent elements is 393.5 kJ or that the enthalpy of formation of carbon dioxide is 393.5 kJ mol-1.

∴ \(6 \mathrm{C}(\mathrm{s})+6 \mathrm{H}_2(\mathrm{~g})+3 \mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{C}_6 \mathrm{H}_{12} \mathrm{O}_6(\mathrm{~s}) \quad \Delta_{\mathrm{f}} H^{\ominus}=-1169 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

When 1 mol of glucose is formed from its constituent elements in their standard states, the heat liberated is 1169 kJ, or the standard molar enthalpy of formation of glucose is 1169 kJ mol-1 (‘molar’ referring to the formation of one mole of the substance).

By convention, the standard enthalpy of formation of elements is zero at all temperatures. However, when an element changes its state, the standard enthalpy of formation is not zero.

Usefulness of \(\Delta_{\mathrm{f}} H^{\ominus}\).

If you know the standard enthalpies of formation of the reactants and products of a chemical reaction, you can easily calculate the standard enthalpy change for that reaction. The difference between the sum of the enthalpies of formation of the products and the sum of the enthalpies of formation of the reactants is the enthalpy change associated with the reaction.

Standard enthalpy = \(\left(\begin{array}{c}
\text { sum of standard heats of } \\
\text { formation of products }
\end{array}\right)-\left(\begin{array}{c}
\text { sum of standard heats of } \\
\text { formation of reactants }
\end{array}\right)\)

∴ \(\Delta_r H^{\ominus}=\Sigma \Delta_f H^{\ominus} \text { (products) }-\Sigma \Delta_f H^{\ominus} \text { (reactants) }\)

Example 1. Calculate the enthalpy change for the following reaction. \(\mathrm{CO}_2(\mathrm{~g})+\mathrm{H}_2(\mathrm{~g}) \longrightarrow \mathrm{CO}(\mathrm{g})+\mathrm{H}_2 \mathrm{O}(\mathrm{g})\)

Given that \(\Delta_t H^{\ominus}\) for CO2(g), CO(g) and H2O(g) are -393.5 kJ mol-1, -111.3 kJ mol-1, and -241.8 kJ mol-1 respectively.

Solution:

⇒ \(\Delta_{\mathrm{r}} H^{\ominus}=\Sigma \Delta_{\mathrm{f}} H^{\ominus} \text { (products) }-\Delta_{\mathrm{f}} H^{\ominus} \text { (reactants) }\)

= \(\left[\Delta_f H^\ominus(\mathrm{CO})+\Delta_{\mathrm{f}} H^{\ominus}\left(\mathrm{H}_2 \mathrm{O}\right)\right]-\left[\Delta_{\mathrm{f}} H^{\ominus}\left(\mathrm{CO}_2\right)+\Delta_{\mathrm{f}} H^{\ominus}\left(\mathrm{H}_2\right)\right]\)

=[(-1113) + (-2418)]-[(-393.5) + (0)] =- 353.1 + 393.5 = 40.4 kJ mol-1.

Example 2. Calculate the enthalpy change for the following reaction.
\(\mathrm{CCl}_4(\mathrm{~g})+2 \mathrm{H}_2 \mathrm{O}(\mathrm{g}) \longrightarrow \mathrm{CO}_2(\mathrm{~g})+4 \mathrm{HCl}(\mathrm{g})_{;} \Delta_{\mathrm{r}} H^{\ominus}=-41.4 \mathrm{kcal}\)

Given that,

⇒ \(\Delta_{\mathrm{f}} \mathrm{H}^{\ominus} \text { for } \mathrm{CCl}_4(\mathrm{~g}), \mathrm{H}_2 \mathrm{O}(\mathrm{g}) \text { and } \mathrm{CO}_2(\mathrm{~g}) \text { are }\) -25.5 k cal mol-1,-57.8 k cal mol-1 and -94.1k cal mol-1 respectively.

Solution:

⇒ \(\Delta_{\mathrm{r}} H^{\ominus} =\left[\Sigma \Delta_{\mathrm{f}} H^{\ominus}(\text { products })\right]-\left[\Sigma \Delta_{\mathrm{f}} H^{\ominus}(\text { reactants })\right]\)

= \(\left[\Delta_{\mathrm{f}} H^{\ominus}\left(\mathrm{CO}_2\right)+4 \times \Delta_{\mathrm{f}} H^{\ominus}(\mathrm{HCl})\right]-\left[\Delta_{\mathrm{f}} H^{\ominus}\left(\mathrm{CCl}_4\right)+2 \times \Delta_{\mathrm{f}} H^{\ominus}\left(\mathrm{H}_2 \mathrm{O}\right)\right]\)

-41.4 = \(\left[(-94.1)+4 \Delta_{\mathrm{f}} H^{\ominus}(\mathrm{HCl})\right]-[(-25.5)+(2 \times-57.8)]\)

or -41.4 = \(-94.1+4 \Delta_{\mathrm{f}} H^{\ominus}(\mathrm{HCl})+141.1\)

∴ \(4 \times \Delta_{\mathrm{f}} H^{\ominus}(\mathrm{HCl})=-41.4-141.1+94.1=-88.4\).

∴ \(\Delta_f H^{\ominus}(\mathrm{HCl})=\frac{-88.4}{4}=-22.1 \mathrm{kcal} \mathrm{mol}^{-1}\).

In our discussion so far, we have not really distinguished between various types of reactions or processes such as combustion, neutralization, and change of state. When we have used the term enthalpy of a reaction, we have used it to refer to any chemical process or reaction. The enthalpies associated with different types of reactions or processes actually go by different names.

Enthalpy of combustion: The enthalpy or heat of combustion of a substance is the heat change accompanying the complete combustion of 1 mol of the substance in (excess) oxygen or air. The heat of combustion of carbon, for instance, is -393.5 kJ mol-1.

⇒ \(\mathrm{C}(\mathrm{s})+\mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{CO}_2(\mathrm{~g}) ; \quad \Delta_{\mathrm{c}} H^{\ominus}=-393.5 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

Standard enthalpy of combustion is defined as the enthalpy per mol of a substance, when all the reactants and products are in their standard states at the specified temperature.

Remember that the heat of combustion of a substance is the heat evolved when 1 mol of the substance is completely burnt or oxidised. Carbon is converted to carbon monoxide according to the following reaction.

⇒ \(\mathrm{C}(\mathrm{s})+\frac{1}{2} \mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{CO}(\mathrm{g}) ; \quad \Delta_c H^{\ominus}=-110.5 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

But -110.5 kJ mol-1 is not the heat of combustion of carbon.

Remember also that combustion reactions are always accompanied by the evolution of heat, so the heat of combustion is always negative.

Calorific value Combustion reactions are very important for us. We bum fuels to meet our energy requirements. We use the energy released by the oxidation of food to survive. The calorific value of a fuel or food is the amount of heat (in calories) released when 1 mol of the fuel or food is burnt or oxidised completely.

Example 1. A cylinder of cooking gas contains 11 kg of butane. The thermochemical equation for the combustion of butane is \(\mathrm{C}_4 \mathrm{H}_{10}(\mathrm{~g})+\frac{13}{2} \mathrm{O}_2(\mathrm{~g}) \longrightarrow 4 \mathrm{CO}_2(\mathrm{~g})+5 \mathrm{H}_2 \mathrm{O}(\mathrm{l}); \quad \Delta_{\mathrm{c}} \mathrm{H}=-2658 \mathrm{~kJ}\)

If a family’s energy requirement per day for cooking is 12000 kJ, how long would a cylinder last?
Solution:

Molar mass of butane = 58 g.

58 g of butane produces = 2658 of heat.

Daily requirement of energy = 12000 kJ.

∴ a cyclinder would last = \(\frac{2658 \times 11 \times 10^3}{58 \times 12,000}=\) = 42 days.

Example 2. Calculate the heat of combustion of glucose from the following data.

  1. \(\mathrm{C}(\mathrm{s})+\mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{CO}_2(\mathrm{~g}); \Delta_{\mathrm{r}} \mathrm{H}_1^{\ominus}=-395.0 \mathrm{~kJ} \mathrm{~mol}^{-1}\)
  2. \(\mathrm{H}_2(\mathrm{~g})+\frac{1}{2} \mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{H}_2 \mathrm{O}(\mathrm{l}) ; \Delta_{\mathrm{r}} \mathrm{H}_2^{\ominus}=-269.4 \mathrm{~kJ} \mathrm{~mol}^{-1}\)
  3. \(6 \mathrm{C}(\mathrm{s})+6 \mathrm{H}_2(\mathrm{~g})+3 \mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{C}_6 \mathrm{H}_{12} \mathrm{O}_6(\mathrm{~s}) ; \Delta_{\mathrm{r}} H_3^{\ominus}=-1169.8 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

Solution: The required equation for the heat of combustion of glucose is

∴ \(\mathrm{C}_6 \mathrm{H}_{12} \mathrm{O}_6+6 \mathrm{O}_2 \longrightarrow 6 \mathrm{CO}_2+6 \mathrm{H}_2 \mathrm{O} \quad \Delta_{\mathrm{r}} \mathrm{H}^{\ominus}=?\)

Applying Hess’s law, the required equation can be obtained by multiplying (1) by 6 and (2) by 6, adding the products and subtracting (3) from the sum.

∴ \(\begin{aligned}
& 6 \mathrm{C}(\mathrm{s})+6 \mathrm{O}_2(\mathrm{~g}) \longrightarrow 6 \mathrm{CO}_2(\mathrm{~g}) \\
& \frac{6 \mathrm{H}_2(\mathrm{~g})+3\mathrm{O}_2(\mathrm{~g})}{} \longrightarrow 6 \mathrm{H}_2 \mathrm{O}(\mathrm{l}) \\
& \hline 6 \mathrm{H}_2(\mathrm{~g})+6 \mathrm{C}(\mathrm{s})+9 \mathrm{O}_2(\mathrm{~g}) \longrightarrow 6 \mathrm{CO}_2(\mathrm{~g})+6 \mathrm{H}_2 \mathrm{O}(\mathrm{l}) \\
&-\left[6 \mathrm{H}_2(\mathrm{~g})+6 \mathrm{C}(\mathrm{s})+3 \mathrm{O}_2(\mathrm{~g})\right.\left.\longrightarrow \mathrm{C}_6 \mathrm{H}_{12} \mathrm{O}_6(\mathrm{~s})\right] \\
& \hline \mathrm{C}_6 \mathrm{H}_1 \mathrm{O}_6(\mathrm{~s})+6 \mathrm{O}_2(\mathrm{~g}) \longrightarrow 6 \mathrm{CO}_2(\mathrm{~g})+6 \mathrm{H}_2 \mathrm{O}(\mathrm{l})
\end{aligned}\)

Then \(\Delta_{\mathrm{r}} H^{\ominus}\) = \(\Delta_{\mathrm{r}} H^{\ominus}=6 \Delta_{\mathrm{r}} H_1^{\ominus}+6 \Delta_{\mathrm{r}} H_2^{\ominus}-\Delta_{\mathrm{r}} H_3^{\ominus}\)

= 6(-395) + 6(-269.4)-(-1169.8)=-2816.6 kJ.

Enthalpy of neutralisation: The enthalpy change associated with the neutralisation of one gram equivalent of an acid by a base (or vice versa) in a dilute aqueous solution is called the enthalpy of neutralisation. For instance, the enthalpy of neutralisation of NaOH by HCl or HCl by NaOH is -57.1 kJ mol-1. In other words, the enthalpy change associated with the neutralisation of one gram equivalent of HCl by NaOH (or vice versa) is -57.1 kJ mol-1.

∴ \(\mathrm{HCl}(\mathrm{aq})+\mathrm{NaOH}(\mathrm{aq}) \longrightarrow \mathrm{NaCl}(\mathrm{aq})+\mathrm{H}_2 \mathrm{O} ; \quad \Delta_{\mathrm{n}} H^{\ominus}=-57.1 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

The enthalpy of neutralisation of any strong acid by a strong base (or vice versa) is always -57.1 kJ mol-1. This is because neutralisation is actually a reaction between the H+ ions produced by an acid and the OH ions produced by a base. Let us see how this is so in the case of HCl and NaOH.

∴ \(\mathrm{Na}^{+}(\mathrm{aq})+\mathrm{OH}^{-}(\mathrm{aq})+\mathrm{H}^{+}(\mathrm{aq})+\mathrm{Cl}^{-}(\mathrm{aq}) \longrightarrow \mathrm{H}_2 \mathrm{O}(\mathrm{l})+\mathrm{Na}^{+}(\mathrm{aq})+\mathrm{Cl}^{-}(\mathrm{aq})\)

or \(\mathrm{H}^{+}+\mathrm{OH}^{-} \longrightarrow \mathrm{H}_2 \mathrm{O}\)

  • When 1 mol of OH ions combines with 1 mol of H+ ions to produce 1 mol of water, the heat liberated is 57.1 kJ mol-1. Strong acids and bases are completely ionised in dilute aqueous solutions and the amount of H+ ions or OH ions produced by one gram equivalent of an acid or a base is always the same, i.e., 1 mol.
  • If either the acid or the base, or both, are weak, the enthalpy of neutralisation is usually less than 57.1 kJ mol-1 in magnitude. This is because weak acids and weak bases do not dissociate completely in an aqueous solution and a part of tire energy liberated during the combination of H+ ions and OH ions is utilised to ionise the weak acid or weak base.
  • The heat used for ionising the weak acid (or base) is called the heat of ionisation or dissociation and the net heat of neutralisation is the difference between 57.1 kJ mol-1 and the heat of dissociation.

∴ \(\underset{\text { Weak acdd }}{\mathrm{CH}_3 \mathrm{COOH}}+\underset{\text { Strong base }}{\mathrm{NaOH}} leftrightharpoons \mathrm{CH}_3 \mathrm{COONa}+\mathrm{H}_2 \mathrm{O} ; \quad \Delta \mathrm{H}=-55.4 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

To calculate the enthalpy of ionisation, let us write down the ionisation reaction and the neutralisation reaction separately.

Ionisation \(\mathrm{CH}_3 \mathrm{COOH} \longrightarrow \mathrm{CH}_3 \mathrm{COO}^{-}+\mathrm{H}^{+} ; \quad \Delta_{\text {ion }} H^{\ominus}=?\)

Neutralisation \(\mathrm{H}^{+}+\mathrm{OH}^{-} \longrightarrow \mathrm{H}_2 \mathrm{O} ; \quad \Delta_{\mathrm{n}} H^{\ominus}=-57.1 \mathrm{~kJ}\)

On adding the two equations, we get, \(\mathrm{CH}_3 \mathrm{COOH}+\mathrm{OH}^{-} \longrightarrow \mathrm{CH}_3 \mathrm{COO}^{-}+\mathrm{H}_2 \mathrm{O}; \Delta_{\text {net }} H^{\ominus}=-55.4 \mathrm{~kJ}\)

The net result is that \(\quad \Delta_{\text {net }} H^{\ominus}=\Delta_{\text {ion }} H^{\ominus}+\Delta_{\mathrm{n}} H^{\ominus} \text { (strong acid-strong base) }\)

or \(\quad-55.4=\Delta_{\text {ion }} H^{\ominus}-57.1 \)

∴\(\quad \Delta_{\text {ion }} H^{\ominus}=-55.4+57.1=2.1 \mathrm{~kJ}\)

In general \(\Delta_{\text {ion }} H^{\ominus}=\left(\Delta_{\text {net }} H^{\ominus}+57.1\right) \mathrm{kJ}\) .

Example: What would be the enthalpy change when

  1. 0.25 mol of HCl in solution is neutralised by 0.25 mol of a NaOH solution?
  2. 0.5 mol of nitric acid in solution is mixed with a solution containing 0.2 mol of a potassium hydroxide solution?
  3. 200 cm3 of a 0.2-M HCl solution is mixed with 300 cm3 of a 0.1-M NaOH solution?

Solution:

1. 0.25 mol of HC1 s 0.25 mol of H+.

0. 25 mol of NaOH = 0.25 mol of OH.

⇒ \(\mathrm{H}^{+}+\mathrm{OH}^{-} \longrightarrow \mathrm{H}_2 \mathrm{O} \quad \Delta_n H^{\ominus}=-57.1 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

Enthalpy change during the formation of 1 mol of H2O = 57.1 kJ.

∴ enthalpy change when 0.25 mol of water is formed = 57.1 x 0.25 = 14.27 kJ.

2. 0.5 mol of HNO3 = 0.5 mol of H+

0. 2 mol of KOH = 0.2 mol of OH

0. 5 mol of H+ ions will react with 0.2 mol of OH ions to produce 0.2 mol of H2O.

Enthalpy change = 0.2 x 57.1 = 1142 kj.

3. 200 cm3 of 0.2-M HCl = 0.04 mol of H+.

300 cm3 of 0.1-M NaOH = 0.03 mol of OH.

0.03 moles of OH will react with 0.04 mol of H+ to produce 0.03 mol of H2O.

Enthalpy change = 57.1 x 0.03 = 1.71 kj.

Enthalpies of phase change: A phase of a system is a homogeneous part of it throughout which all physical and chemical properties are the same. You will study phases and phase changes in detail later. A couple of examples will give you a basic idea.

  • If you have a beaker full of water, the system you are considering consists only of liquid water, so it has one phase. If you drop a few cubes of ice in the water, your system now contains water and ice, so it is a two-phase system.
  • When a system changes from one phase to another the process is referred to as a phase change or phase transition. Thus, the conversion of a solid into a liquid (fusion) is a phase change, as also the conversion of a liquid into a gas (vaporisation).
  • There are other kinds of phase changes, but we are not concerned with those here. You already know that energy is needed to change a solid to a liquid and a liquid to a gas. The enthalpy changes associated with phase transitions go by different names. Remember, during a phase change, the temperature remains constant.

Enthalpy of fusion The enthalpy of fusion of a substance is the enthalpy change accompanying the conversion of 1 mol of the substance in its solid state into its liquid state at its melting point. For example, the standard enthalpy of fusion of ice is 6.0 kJ mol-1.

∴ \(\mathrm{H}_2 \mathrm{O}(\mathrm{s}) \longrightarrow \mathrm{H}_2 \mathrm{O}(\mathrm{l}) ; \quad \Delta_{\text {fus }} H^{\ominus}=+6.0 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

∴ \(\Delta_{\text {fus }} H^{\ominus}\) is the enthalpy of fusion in the standard state. Freezing is the reverse process of fusion, in that case an equal amount of heat is given off to the surroundings.

∴ \(\Delta_{\text {freex }} H^{\ominus}=-\Delta_{\text {fus }} H^{\ominus}\)

Enthalpy of vaporisation The enthalpy of vaporisation of a substance is the enthalpy change accompanying the conversion of 1 mol of the substance in its liquid state into its vapour state at the boiling point of the liquid. The standard enthalpy of vaporisation of water is 40.7 kj mol-1.

∴ \(\mathrm{H}_2 \mathrm{O}(\mathrm{l}) \longrightarrow \mathrm{H}_2 \mathrm{O}(\mathrm{g}) ; \quad \Delta_{\text {vap }} H^{\ominus}=40.7 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

Enthalpy of sublimation The enthalpy of sublimation of a substance is the heat change associated with the conversion of 1 mol of it directly from its solid to its gaseous state at a temperature below its melting point. For instance, the standard enthalpy of sublimation of iodine is 62.39 kJ mol-1.

∴ \(\mathrm{I}_2(\mathrm{~s}) \longrightarrow \mathrm{I}_2(\mathrm{~g}) ; \quad \Delta_{\mathrm{sub}} H^{\ominus}=61.4 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

The enthalpy change of n reverse transition is the negative of the enthalpy change of the forward transition (under the same conditions). This further proves enthalpy to be a state property. Therefore, the change in enthalpy would be same in both the cases when a solid is directly or indirectly converted to vapour. Thus, the enthalpy of sublimation can be expressed as:

Basic Chemistry Class 11 Chapter 6 Thermodynamics Freezing Points And Standard Enthalpies Of physical Change Of Some Common Compounds

⇒ \(\Delta_{\text {sub }} H^{\ominus}=\Delta_{\text {fus }} H^\ominus+\Delta_{\text {vap }} H^{\ominus}\)

  • The striking differences in the magnitude of the enthalpy change for various substances is attributed to the intermolecular interaction in the substances. The enthalpy of vaporisation of water at its boiling point is 40.79 kJ mol-1.
  • This signifies that water molecules are held together more tightly than any other liquid with low enthalpy of vaporisation, for instance, acetone. The high enthalpy of vaporisation of water is partly responsible for low humidity in the atmosphere.

Enthalpy of atomisation: The enthalpy change on breaking a molecule completely into its gaseous atoms is called enthalpy of atomisation \(\Delta_{\mathrm{a}} H\). In case of a metal, the enthalpy of atomisation is tine same as the enthalpy of sublimation.

⇒ \(\mathrm{C}(\mathrm{s}) \longrightarrow \mathrm{C}(\mathrm{g}) \Delta_{\mathrm{a}} H^{\ominus}=716.7 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

⇒ \(\mathrm{Na}(\mathrm{s}) \longrightarrow \mathrm{Na}(\mathrm{g}) \Delta_{\mathrm{a}} H^{\ominus}=108.4 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

The enthalpy of atomisation of diatomic gases is half the bond dissociation enthalpy. For example, the standard bond dissociation enthalpy of O2 is 497 kJ mol-1 and the standard enthalpy of atomisation pertaining to the reaction is 249.2 kJ mol-1.

⇒ \(\frac{1}{2} \mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{O}(\mathrm{g})\)

Enthalpy of allotropic transformation: You know that elements like carbon, sulfur, and phosphorus can exist in different allotropic forms and that such elements can change from one allotropic form into another. Allotropic transformations involve enthalpy changes.

  • The enthalpy of the allotropic transformation of one allotropic form of a substance into another is the heat change accompanying the transformation per mole of the substance. The standard enthalpy of transformation of rhombic sulphur to monoclinic sulfur, for example, is 1.3 kJ mol-1.
  • It is not easy to determine enthalpy changes for allotropic transformations experimentally because such processes are rather slow and the enthalpy changes associated with them are small. Hess’s law can be used conveniently to determine the enthalpy changes accompanying allotropic transformations.

Example: Calculate the enthalpy of the allotropic transformation of rhombic sulphur to monoclinic sulphur from the following thermochemical equations.

  1. \(
    \mathrm{S}\left(\text { rhombic) }+\mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{SO}_2(\mathrm{~g})\right. \Delta_{\mathrm{c}} H_1^{\ominus}=-295.1 \mathrm{~kJ} \mathrm{~mol}^{-1}\) …(1)
  2. \(\mathrm{~S}(\text { monoclinic })+\mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{SO}_2(\mathrm{~g}) \Delta_{\mathrm{c}} H_2^{\ominus}=-296.4 \mathrm{~kJ} \mathrm{~mol}^{-1}\)……(2)

Solution:

The required equation is S(rhombic) → S(monoclinic) ΔH =?

Applying Hess’s law and subtracting equation (2) from (1)

∴ \(\begin{array}{rc}
\mathrm{S}(\mathrm{r})+\mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{SO}_2(\mathrm{~g}) & \Delta_{\mathrm{c}} H_1=-295.1 \mathrm{~kJ} \\
\mathrm{~S}(\mathrm{~m}) \pm \mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{SO}_2(\mathrm{~g}) & \Delta_{\mathrm{c}} H_2=-296.4 \mathrm{~kJ} \\
\hline \mathrm{S}(\mathrm{r}) \longrightarrow \mathrm{S}(\mathrm{m}) & \Delta H=13 \mathrm{~kJ}
\end{array}\)

Therefore, the enthalpy of the allotropic transformation of rhombic sulphur to monoclinic sulphur is 1.3 kJ mol-1.

Enthalpy of solution: The enthalpy of reaction for dissolving one mole of a solute in n moles of a solvent is known as the integral enthalpy of the solution or the integral heat of the solution. Let us look at some such values for a solution of HNO3 in water at 298 K.

1. \(\left.\mathrm{HNO}_3(\mathrm{l})+\mathrm{H}_2 \mathrm{O} \longrightarrow \mathrm{HNO}_3 \text { (in } 1 \mathrm{H}_2 \mathrm{O}\right) \Delta_{\text {sol }} H^{\ominus}=-187.6 \mathrm{~kJ} \mathrm{~mol}^{-1}\) ….(1)

2. \(\left.\mathrm{HNO}_3(\mathrm{l})+10 \mathrm{H}_2 \mathrm{O} \longrightarrow \mathrm{HNO}_3 \text { (in } 10 \mathrm{H}_2 \mathrm{O}\right) \Delta_{\text {sol }} H^{\ominus}=-205.8 \mathrm{~kJ} \mathrm{~mol}^{-1}\) …. (2)

3. \(\mathrm{HNO}_3(\mathrm{l})+100 \mathrm{H}_2 \mathrm{O} \longrightarrow \mathrm{HNO}_3 \text { (in } 100 \mathrm{H}_2 \mathrm{O} \text { ) } \Delta_{\text {sol }} H^{\ominus}=-206.8 \mathrm{~kJ} \mathrm{~mol}^{-1}\)…. (3)

4. \(\mathrm{HNO}_3(\mathrm{l})+500 \mathrm{H}_2 \mathrm{O} \longrightarrow \mathrm{HNO}_3 \text { (in } 500 \mathrm{H}_2 \mathrm{O} \text { ) } \Delta_{\text {sol }} H^{\ominus}=-206.9 \mathrm{~kJ} \mathrm{~mol}^{-1}\)…..(4)

5. \(\left.\mathrm{HNO}_3(\mathrm{l})+1000 \mathrm{H}_2 \mathrm{O} \longrightarrow \mathrm{HNO}_3 \text { (in } 1000 \mathrm{H}_2 \mathrm{O}\right) \Delta_{\text {sol }} H^{\ominus}=-206.9 \mathrm{~kJ} \mathrm{~mol}^{-1}\)….(5)

6. \(\mathrm{HNO}_3(\mathrm{l})+\mathrm{aq} \longrightarrow \mathrm{HNO}_3 \text { (aq) } \Delta_{\text {sol }} H^{\ominus}=-207.4 \mathrm{~kJ} \mathrm{~mol}^{-1}\) …..(6)

As you can see from Equations (1)-(6), the heat of solution varies with concentration or decreases with increase in the amount of the solvent and finally stabilises at infinite dilution.

The enthalpy of solution at infinite dilution is the enthalpy change when 1 mol of the substance dissolves in such a large amount of solvent that the interaction between the solute molecules is negligible. In case of water as a solvent, infinite dilution is denoted by ‘aq’ (aqueous).

For HNO3, the enthalpy of solution at infinite dilution refers to Equation (6) and is -207.36 kJ mol-1.

Some more values of enthalpy of solution at infinite dilution are given below.

⇒ \(\mathrm{KCl}(\mathrm{s})+\mathrm{aq} \longrightarrow \mathrm{KCl}(\mathrm{aq}) ; \Delta_{\mathrm{sol}} H^{\ominus}=+18.6 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

⇒ \(\mathrm{CuSO}_4(\mathrm{~s})+\mathrm{aq} \longrightarrow \mathrm{CuSO}_4(\mathrm{aq}) ;  \Delta_{\mathrm{sol}} H^{\ominus}=-66.5 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

⇒ \(\mathrm{CuSO}_4 \cdot 5 \mathrm{H}_2 \mathrm{O}+\mathrm{aq} \longrightarrow \mathrm{CuSO}_4(\mathrm{aq}) ; \Delta_{\text {sol}} H^{\ominus}=+11.7 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

  • In the first and third cases \(\Delta_{\text {sol}}\) is positive (endothermic), while in the second, it is negative (exothermic). In general, salts which do not form hydrates dissolve in water with the absorption of heat. Hydrated salts also dissolve in water with the absorption of heat. The process is exothermic only in the case of anhydrous salts which form hydrates.
  • When a substance dissolves in water, it either absorbs (positive enthalpy) or releases energy (negative enthalpy). Ammonium nitrate has a positive enthalpy of solution. Tire cold pack used for minor injuries to sportsmen contains NH4NO3 and water in separate compartments.
  • It is activated on punching the partition between the two upon which the salt dissolves and the temperature of water falls. Calcium chloride, on the other hand, has a negative enthalpy of solution, and this phenomenon is used to dissolve ice on sidewalks in winters.
  • Let us see what happens when NaCl dissolves in water. First, the three-dimensional network of Na+ and Cl breaks into individual ions. For this purpose the lattice energy which is responsible for holding the ions together must be overcome. The separated Na+ and Cl ions get stabilised in tire solution by their interaction with the water molecules.

This process involves breaking some of the hydrogen bonds between the water molecules. The ions are then said to be hydrated. The process of dissolution of an ionic compound in a solvent (water) involves a complex interaction between the solute and the solvent species. However, it can be broadly taken to be happening in two steps:

1. First the ionic lattice breaks: \(\mathrm{NaCl}(\mathrm{s}) \longrightarrow \mathrm{Na}^{+}(\mathrm{g})+\mathrm{Cl}^{-}(\mathrm{g})\)

Energy is required for this process and therefore the reaction is endothermic. The enthalpy change associated with it is called lattice enthalpy it may be defined as the heat change when one mole of an ionic solid separates into its constituent gaseous atoms.

2. Next the gaseous Na+ and Cl ions interact with water molecules and become hydrated:

⇒ \(\mathrm{Na}^{+}(\mathrm{g})+\mathrm{Cl}^{-}(\mathrm{g}) \stackrel{\mathrm{H}_2 \mathrm{O}}{\longrightarrow} \mathrm{Na}^{+}(\mathrm{aq})+\mathrm{Cl}^{-}(\mathrm{aq})\)

The enthalpy change associated with this process is called the enthalpy of solvation or more precisely the enthalpy of hydration as water is the solvent.

∴ \(\Delta_{\text {sol }} H=\Delta_{\text {lattlee }} H+\Delta_{\text {hyd }} H\)

Lattice enthalpy: It is not possible to determine lattice energy directly by experiments. Hence, it is determined on the basis of the Born-Haber cycle. The Bom-Haber cycle is a closed sequence of different processes which are involved in the breaking and finally in the making of an ionic crystal. To understand the cycle, let us take the example of NaCl(s). The steps involved in the formation of NaCl(s) are as follows:

1. Sublimation of Na(s) to Na(g): \(\mathrm{Na}(\mathrm{s}) \longrightarrow \mathrm{Na}(\mathrm{g}) \quad \Delta H=\Delta_{\mathrm{a}} H^{\ominus}[\mathrm{Na}(\mathrm{s})]=108.4 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

2. Dissociation of Cl2(g) to give Cl(g): \(\frac{1}{2} \mathrm{Cl}_2(\mathrm{~g}) \longrightarrow \mathrm{Cl}(\mathrm{g}) \quad \Delta H=\frac{1}{2} \Delta_{\mathrm{Cl}-\mathrm{C}} H^{\ominus}=\frac{1}{2} \times 242 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

3. Ionisation of Na(g): \(\mathrm{Na}(\mathrm{g}) \longrightarrow \mathrm{Na}^{+}(\mathrm{g})+\mathrm{e}^{-} \quad \Delta H=\Delta_{\text {ien }} H^{\ominus}=495.8 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

Ionisation enthalpy of sodium = 495.8 kJ mol-1. Since sodium is in the gaseous state, the enthalpy change is given by

∴ \(\Delta_{\text {ion }} H=\Delta_{\text {ion }} U+\Delta n_g R T\)

∴ \(\Delta_{\text {ion }} H=495.8+(1) \times R T\)

(It may be observed subsequently that RT cancels out with -RT in the next step.)

4. The electron released by the sodium atom is gained by chlorine to give Cl.

⇒ \(\mathrm{Cl}(\mathrm{g})+\mathrm{e}^{-} \longrightarrow \mathrm{Cl}^{-}(\mathrm{g})\)

The electron affinity (EA) of Cl is 348.6 kJ mol-1. Since electron affinity is defined as the energy released in the above process, from thermodynamic considerations we must take a negative value.

Hence energy involved (-EA) =-348.6 kJ mol-1. Here again, we need the electron gain enthalpy (\(\Delta_{\mathrm{eg}} H)\), which is -EA – RT (since Δng = -1)

∴ \(\Delta_{\mathrm{eg}} H)\) = -348.6 + (-1)RT = -348.6 – RT.

5. Na+(g) and Cl(g) (steps 3 and 4) combine to give NaCl(s).

⇒ \(\mathrm{Na}^{+}(\mathrm{g})+\mathrm{Cl}^{-}(\mathrm{g}) \longrightarrow \mathrm{NaCl}(\mathrm{s})\)

Enthalpy change for this process = –\(\Delta_{\text {Laltice }} H\)(The process involves the formation of the lattice.) These steps are depicted as the sequence of steps in the Bom-Haber cycle.

Basic Chemistry Class 11 Chapter 6 Thermodynamics Born Haber Cycle Of NaCl

The basis of calculations using the above cycle is that enthalpy being a state function, the sum of enthalpy changes around the cycle (clockwise direction) is zero.

∴ \(-\Delta_{\mathrm{f}} H[\mathrm{NaCl}(\mathrm{s})]+\Delta_{\text {sub }} H[\mathrm{Na}(\mathrm{s})]\)+\(\frac{1}{2} \Delta_{\text {diss }} H\left[\mathrm{Cl}_2(\mathrm{~g})\right]+I \cdot E .+R T-E . A-R T-\Delta_{\text {lattice }}\) H =0

or \(\Delta_{\text {lattice }} H=-\Delta_{\mathrm{f}} H+\Delta_{\text {sub }} H+\frac{1}{2} \Delta_{\text {diss }} H+I \cdot E .-E . A .\)

The \(\Delta_{\mathrm{f}}H\) of NaCl(s) is -411.2 kJ.

Substituting the appropriate values on the RHS, we get

∴ \(\Delta_{\text {Lattice }} H =411.2+108.4+\frac{1}{2} \times 242+495.8-348.6\) =1136.4-348.6

∴ \(\Delta_{\text {Lattice }} H =787.8 \mathrm{~kJ}\).

Bond energy: By now you know what enthalpies of formation of compounds are. You also know that compounds are formed by the breaking and making of bonds. To have a precise knowledge of the enthalpy change for a process, the bond dissociation enthalpy or the bond dissociation energy [ΔH(A-B)] is considered.

For example, when the dissociation, or breaking, of a chemical bond occurs as in the following process,the corresponding molar enthalpy change is called the bond dissociation enthalpy. Thus, the bond dissociation enthalpy is tire enthalpy associated with the breaking of 1 mol of a substance in the gaseous state completely into its gaseous atoms.

∴ AB(g) → A(g) + B(g)

The values of bond enthalpies depend on the bonding present between two atoms in the molecule. Even in the same molecule, the values may differ. For example, to break the first O—H bond in water, the enthalpy change \(\Delta H^{\ominus}(\mathrm{HO}-\mathrm{H})\)= 492 kJ mol-1 while for breaking the second bond, \(\Delta H^{\ominus}(\mathrm{O}-\mathrm{H})\) is 428 kJ mol-1.

  • In methane four hydrogen atoms are attached to a carbon atom and the four C—H bonds are equal in energy and bond length. In this case the total enthalpy is expected to be four times the C—H bond enthalpy.
  • However, this is not the case, the reason being different dissociation steps. The first step involves the breaking of a C—H bond in the CH4 molecule but in the next step the C—H bond is broken in a CH3 radical, and so on.

The energies required to break the individual C—H bonds in different dissociation steps are as follows.

∴ \(\mathrm{CH}_4(\mathrm{~g}) \longrightarrow \mathrm{CH}_3(\mathrm{~g})+\mathrm{H}(\mathrm{g}) ; \Delta_{\text {bond }} H^{\ominus}=427 \mathrm{~kJ} \mathrm{~mol}^{-1},\)

∴ \(\mathrm{CH}_3(\mathrm{~g}) \longrightarrow \mathrm{CH}_2(\mathrm{~g})+\mathrm{H}(\mathrm{g}) ; \Delta_{\text {bond }} H^{\ominus}=439 \mathrm{~kJ} \mathrm{~mol}^{-1},\)

∴ \(\mathrm{CH}_2(\mathrm{~g}) \longrightarrow \mathrm{CH}(\mathrm{g})+\mathrm{H}(\mathrm{g}) ; \Delta_{\text {bond }} H^{\ominus}=452 \mathrm{~kJ} \mathrm{~mol}^{-1},\)

∴ \(\mathrm{CH}(\mathrm{g}) \longrightarrow \mathrm{C}(\mathrm{g})+\mathrm{H}(\mathrm{g}) ; \Delta_{\text {bond }} H^{\ominus}=347 \mathrm{~kJ} \mathrm{~mol}^{-1}\).

The total enthalpy change for the atomisation of methane \(\left[\mathrm{CH}_4(\mathrm{~g}) \longrightarrow \mathrm{C}(\mathrm{g})+4 \mathrm{H}(\mathrm{g})\right] \text { is } \Delta_{\mathrm{a}} H^{\ominus}\) = 166 x 103 kJ mol-1.

In case of reactions for which experimental data are not available for successive steps, mean bond enthalpy is used. It is the value of bond dissociation energy of a bond A—B averaged over a series of related compounds. For example, the bond energy for the O—H bond, \(\Delta H^{\ominus}\)(O—H) (averaged between H2O and alcohols) is 463 kJ mol-1.

Similarly, in case of CH4, all the bond enthalpies added together come to 1665 kJ mol-1. The mean bond enthalpy of C—H bond in CH4 is

⇒ \(\frac{1}{4} \Delta_a H^{\ominus}=\frac{1}{4} \times 1665=416 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

Now consider the atomisation of a diatomic molecule: \(\mathrm{H}_2(\mathrm{~g}) \longrightarrow 2 \mathrm{H}(\mathrm{g}) ; \quad \Delta_a H^{\ominus}=435 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

Since two atoms are formed, the energy value is the enthalpy of atomisation of H2 at 298 K.

The corresponding bond energy \(\Delta_a U^{\ominus}\) for the same reaction (with a little difference from the enthalpy) is 430.8 kJ mol-1. That is why the two quantities are often used interchangeably. But the correct way would be to convert from ΔU to ΔH.

We know that ΔH = ΔU + pΔV

Some important bond enthalpies (kJ mol-1) of multiple bonds are as follows.

C—C = 612 N=N = 418 O=O= 497

C≡C = 837 N≡N = 946

C=O = 741 C=N = 615

  • For gases, pΔV (at constant pressure) may be replaced by RT. In case of a diatomic molecule, the bond energy is equal to the bond dissociation energy.
  • The dissociation energy may here be defined as the enthalpy change involved in breaking the bond between atoms of a gaseous diatomic molecule. Also, it can be seen that bond enthalpy in homonuclear diatomic molecules is twice the enthalpy of formation for the atom in the gaseous state.

Bond energies can be used to calculate the enthalpy of a reaction. The standard reaction enthalpy, \(\Delta_r H^{\ominus}\) is the difference between the sum of the standard enthalpies of the reactants and products.

∴ \(\Delta_r H^{\ominus}=\Sigma \text { bond enthalpies }{ }_{\text {reactants }}-\Sigma \text { bond enthalpies products }\)

This relationship is valid when all the substances in a chemical reaction are in the gaseous state.

Example 1. Calculate the enthalpy change for the following reaction. \(2 \mathrm{C}_2 \mathrm{H}_2(\mathrm{~g})+5 \mathrm{O}_2(\mathrm{~g}) \longrightarrow 4 \mathrm{CO}_2(\mathrm{~g})+2 \mathrm{H}_2 \mathrm{O}(\mathrm{g})\)

The average bond energies of the C—H, C≡C, O=O, C—O, and O—H bonds are 414, Hit), 499, 724, and 460 kJ mol-1 respectively.
Solution:

ΔH = sum of bond energies of reactants – sum of bond energies of products.

The given reaction can be written as 2H—C≡C—H(g) + 5O=O(g) → 4O=C=)(g) + 2H—O—H(g)

∴ \(\Delta H=\left[2 \Delta H_{\mathrm{C} \equiv \mathrm{C}}+4 \Delta H_{\mathrm{C}-\mathrm{H}}+5 \Delta H_{\mathrm{O}=\mathrm{O}}\right]-\left[8 \Delta H_{\mathrm{C}=\mathrm{O}}+4 \Delta H_{O-H}\right]\)

= (2 x 810 + 4 x 414 + 5 x 499) – (8 x 724 + 4 x 460)

= 5771 – 7632 = -1861 kJ mol-1.

Example 2. Calculate the bond energy of the C—H bond in CH4 front the following data.

  1. \(\mathrm{C}(\mathrm{s})+2 \mathrm{H}_2(\mathrm{~g}) \longrightarrow \mathrm{CH}_4(\mathrm{~g}) \quad \Delta H_1^{\ominus}=-74.8 \mathrm{~kJ} \mathrm{~mol}^{-1}\) ….(1)
  2. \(\mathrm{H}_2(\mathrm{~g}) \longrightarrow 2 \mathrm{H}(\mathrm{g}) \quad \Delta \mathrm{H}_2^{\ominus}=435.4 \mathrm{~kJ} \mathrm{~mol}^{-1}\) …..(2)
  3. \(\mathrm{C}(\mathrm{s}) \longrightarrow \mathrm{C}(\mathrm{g}) \quad \Delta H_3^{\ominus}=718.4 \mathrm{~kJ} \mathrm{~mol}^{-1}\)….(3)

Solution:

To find the bond energy of the C—H bond in CH4, let us first write the equation for the dissociation of CH4 into C and H, The enthalpy change for this reaction would be four times the bond energy of the C—H bond since the dissociation of the CH4 molecule involves the breaking of four C—H bonds.

CH4 (g) → C(g) + 4H(g) ΔH = ?

The enthalpy change for this reaction can be calculated from the given data by applying Hess’s law. The equation for this reaction can be obtained by multiplying Equation (2) by 2, adding the product to Equation (3), and then subtracting Equation (1) from the sum.

∴ \(\begin{aligned}
\mathrm{C}(\mathrm{s})+2 \mathrm{H}_2(\mathrm{~g}) & \longrightarrow 4 \mathrm{H}(\mathrm{g})+\mathrm{C}(\mathrm{g}) \\
\mathrm{C}(\mathrm{s})+2 \mathrm{H}_2(\mathrm{~g}) & \longrightarrow \mathrm{CH}_4(\mathrm{~g}) \\
\hline \mathrm{CH}_4(\mathrm{~g}) & \longrightarrow 4 \mathrm{H}(\mathrm{g})+\mathrm{C}(\mathrm{g})
\end{aligned}\)

∴ ΔH = 2 x ΔH2 + ΔH3 – ΔH1

= 2 x 435.4 + 718.4-(-748)

= 870.8 + 718.4 + 748 = 16640 kJ mol-1.

Thus, the energy required to break four C—H bonds = 1664.0 kJ mol-1.

Therefore, the bond energy of the C—H bond = 1664/4 = 416 kJ mol-1.

Alternatively, the enthalpy of the formation of the reactants and products can also be used to find the enthalpy change for the required reaction.

ΔH = ∑ΔHf (products) -∑ΔHf (reactants)

ΔH = [ΔH3 +2ΔH2]-ΔH1

= [718.4 + 2 x 435.4] – [-748]

= 718.4 + 870.4 + 74.8 = 1664.0 kJ mol-1,

∴bond energy of C—H bond = 1664/4 = 416 kJ mol-1.

The Measurement Of Heat

In the laboratory, heat changes in physical and chemical processes are measured with a calorimeter. Calorimetry is the measurement of heat changes associated with a process. The study of calorimetry involves the concept of specific heat and heat capacity. So let us define these first

The heal capacity (C) of a substance is the amount of heat required to raise the temperature of a given quantity of it by 1°C Heat capacity is an extensive property. It is more convenient to use an intensive property. Therefore, the molar heat capacity, Cm (Cm = C/n in JK-1 mol-1 where n is the amount of the substance), is used. The specific heat capacity Cs (Cs = C/m in JK-1 g-1, where m is the mass of the substance) is also an intensive property.

For water, the specific heat is \(1 \mathrm{cal} \mathrm{g}^{-1} \mathrm{C}^{-1} \text { or } 4.18 \mathrm{~J} \mathrm{~g}^{-10} \mathrm{C}^{-1}\)

The molar heat capacity of a substance is the heat required to raise the temperature of 1 mol of the substance by one degree.

The heat capacity at constant volume (Cv) is given by \(C_v=\frac{q_v}{\Delta T} .\)

From the first law of thermodynamics, we have q = ΔU ÷ pΔV.

At constant volume, ΔV = 0. qv = AU.

∴ heat capacity at constant volume is given by \(C_V=\left(\frac{\Delta U}{\Delta T}\right)_V\)

Thus, the heat capacity at constant volume is defined as the rate of change of internal energy with temperature.

The heat capacity at constant pressure is given by \(C_p=\frac{q_p}{\Delta T}.\)

We have already learned that, at constant pressure, ΔH = qp.

Therefore, heat capacity at constant pressure is given by \(C_p=\left(\frac{\Delta H}{\Delta T}\right)_p\)

Thus, the heat capacity at constant pressure is defined as the rate of change of enthalpy with temperature. Relationship between Cp and Cv We know that, at constant pressure (Equation 6),

ΔH = ΔU + pΔV.

For 1 mole of an ideal gas, pΔV = RΔT (ideal gas equation).

Thus, Equation 6 can be written as ΔH = ΔU + RΔT

On dividing the equation by ΔT, we get \(\frac{\Delta H}{\Delta T}=\frac{\Delta U}{\Delta T}+R\)

or Cp = Cv + R

or Cp – Cv = R

Cp is always greater than Cv for any gas, since in a constant-pressure process a system has to do work against the surroundings.

Coming back to the measurement of energy changes associated with chemical or physical processes, measurements are made under two different conditions:

  1. at constant volume, ΔU -qv, and
  2. at constant pressure, ΔH = qp.

Constant-volume calorimetry: The reactions that can be most easily studied under constant- volume conditions are combustion reactions. The apparatus used for this purpose is called the bomb calorimeter or adiabatic bomb calorimeter. It consists of a sealed constant-volume steel container (known as a bomb) in which the reaction occurs. The bomb can withstand high pressures.

  • A known mass of a combustible substance is placed in the bomb, which is filled with oxygen at 30 atm pressure. The sealed bomb is then immersed in a known amount of water contained in an insulated container. The whole set-up is called the calorimeter. Since the outer covering is insulated, there is no heat exchange with the surroundings and hence the reaction process is adiabatic.
  • The sample is ignited electrically and the heat produced by the combustion reaction is absorbed by the water surrounding the bomb and the bomb itself. The temperature of the water is monitored and the heat produced is calculated using the equation

∴ \(q=C_{\text {cal }} \Delta T,\) ….(1)

where Ccal is the heat capacity of the calorimeter. Ccal is determined separately by igniting a known amount of a substance whose heat of combustion is already known, in the same calorimeter.

  • Thus the heat produced in the reaction may be determined. We already know that at constant volume qv = ΔU. In constant- volume calorimetry, the heat change observed is simply the change in internal energy.
  • This value may be modified to get the enthalpy, but the pressure changes are usually small and may be neglected. The heat changes may be simply assumed to correspond to enthalpy changes.

Basic Chemistry Class 11 Chapter 6 Thermodynamics A Constant Volume Bomb Calorimeter

Constant-pressure calorimetry: To measure the heat changes at constant pressure, a device comparable to a thermos flask is used. The calorimeter is just a vessel with a stirrer and a thermometer.

The system is insulated from the surroundings and hence acts as an isolated system. The heat changes within the vessel are measured in terms of the change in temperature of the system. Constant monitoring of the temperature before and after the reaction helps in finding q and hence AH, as we know that, ΔH = qp

The amount of heat exchanged (qv or qp) in any process is given by, qp =msΔT……(2)

where m is the mass of the substance, s is its specific heat and ΔT is the change in the temperature.

You already know that q = CΔT =Cp ΔT (at constant pressure),

where C = ms, the heat capacity of the system.

  • Thus to determine q, C and ΔT have to be known first. The experimental technique—calorimetry—actually involves two steps, the first being the determination of the heat capacity of the calorimeter.
  • The second step involves the determination of the change in temperature during the completion of the reaction. Since it is an insulated system, no heat is lost to the surroundings as in the case of constant-volume calorimetry.

As you already know, the enthalpy of the reaction is simply the heat change at constant pressure,  ΔH = qp (at constant pressure).

Basic Chemistry Class 11 Chapter 6 Thermodynamics A Constant Pressure Calorimeter

Enthalpy changes are commonly tabulated at 25°C and it is often required to know their values at other temperatures. They can be calculated from the heat capacities of the reactants and products.

You know that ΔH = H(products) – H(reactants).

At constant pressure ΔH /ΔT = Cp.

Therefore, for small changes in temperature, \(\frac{\Delta(\Delta H)}{\Delta T}=C_p \text { (products) }-C_p \text { (reactants) }=\Delta C_p \Delta T\)

∴ \(\Delta(\Delta H)=\Delta C_p\left(T_2-T_1\right)\)

or \(\Delta H_2-\Delta H_1=\Delta C_p\left(T_2-T_1\right)\),

where ΔH2 is the enthalpy of the reaction at T2 and ΔH1 is that at T1.

Example 1. Calculate the enthalpy change on the freezing of 1 mol of zoater at -5°C to ice at -5°C. Given that

\(\Delta_{\text {tom }} H^{\oplus}\left(\mathrm{H}_2 \mathrm{O}\right)=6.03 \mathrm{~kJ} \mathrm{~mol}^{-1} \text { at } 0^{\circ} \mathrm{C} \text { and } C_p\left[\mathrm{H}_2 \mathrm{O}(\mathrm{l})\right]\)

= \(75.3 \mathrm{~J} \mathrm{~K}^{-1} \mathrm{~mol}^{-1}, C_{,},\left[\mathrm{H}_2 \mathrm{O}(\mathrm{s})\right]=36.8 \mathrm{~J} \mathrm{~K}^{-1} \mathrm{~mol}^{-1} \text {. }\)

Solution:

To find the enthalpy change at -5°C, we use the equation: \(\Delta H_2-\Delta H_1=\Delta C_p\left(T_2-T_1\right) .\)

First, we will find Δfus H at -5°C and then convert it to ΔH (freezing) by prefixing a (-) sign to ΔfusH.

Also, ΔH is in kJ and Cp in J. Converting, ΔfusH = 6.03 kJ mol-1 = 6.03 x 1000 J mol-1.

∴ \(\Delta H_2-6.03 \times 1000= (75.3-36.8)\times(-5-0) \)

= \(C_p(1)-C_p(\mathrm{~s})=-192.5\)

∴ \(\Delta H_2=-192.5+6030=5837.5\)

Thus \(\Delta_{\text {fus }} H at -5^{\circ} \mathrm{C}=5837.5 \mathrm{~J} \mathrm{~mol}^{-1}=5.84 \mathrm{~kJ} \mathrm{~mol}^{-1} and \Delta H (freezing) at -5^{\circ} \mathrm{C}\)

= \(-5.84 \mathrm{~kJ} \mathrm{~mol}^{-1}\).

Example 2. Find the heat required to raise the temperature of 27.9 g of Fe from 30 to 40°C. The molar heat capacity of Fe(s) =25.10 J K-1mol-1.
Solution:

The molar heat capacity is the heat required to raise the heat of 1 mol of the substance, i.e., 55.8 g (molar mass of Fe) by 1°C.

The heat required to raise the temperature of 55.8 g of Fe by 1°C = 25.1 J.

The heat required to raise the temperature of 27.9 g of Fe by 1°C = \(\frac{25.1}{55.8}\) x 27.9 J.

∴ the heat required to raise the temperature of 27.9 g of Fe by 10°C = \(\frac{25.1}{55.1}\)x 27.9 x 10 = 125.5 J.

Example 3. 1.922 g of methanol (CH3OH) was burnt in a constant-volume bomb calorimeter. The temperature of water rose by 4.2°C. If the heat capacity of the calorimeter and its contents is 10.4 kJ °C-1, calculate the molar heat of combustion of methanol.
Solution:

If qcal is the quantity of heat involved in the reaction and C„ is the heat capacity of the calorimeter, then

∴ \(q_{\text {eat }}=C_V \Delta T\)

But \(q_{\text {cal }}=-q_r\)

∴ \(\quad q_r=-C_v \Delta T\)

= \(-10.4 \times 4.2\left[because \text { the temperature rose by } 4.2^{\circ} \mathrm{C}, \Delta T \text { is positive }\right]\) = \(-43.68 \mathrm{~kJ}\).

The minus sign indicates exothermic nature of the reaction.

Since the pressure changes are small and can be neglected, qr = ΔHr.

Tim ΔH for burning 1.922 g of methanol is -43.68 kJ.

The ΔH for burning 1 mol or 32 g of methanol is obtained as

∴ \(\Delta H=\frac{-43.68 \mathrm{~kJ}}{1.922 \mathrm{~g}} \times 32 \mathrm{~g} \mathrm{~mol}^{-1}\) = -727.24 kJ mol-1.

The molar enthalpy or molar heat of combustion is therefore -727.24 kJ mol-1.

Spontaneous Processes

Generally speaking, when we say something has happened spontaneously, we mean that it has happened of its own accord. In chemistry, the meaning is a little different. It would be easy enough for you to associate the word spontaneous with a process like the dissolving of sugar in water, or the evaporation of water from an open vessel.

  1. But the burning of coal in air, or the burning of ethane to produce carbon dioxide and water are also spontaneous processes, even though these processes have to be initiated by ignition.
  2. Any process which can occur under a given set of conditions is called a spontaneous process, irrespective of whether it occurs on its own, or needs to be initiated. In other words, a spontaneous process occurs on its own or has a tendency to occur. A simple way of saying this is that a spontaneous process is a process which is possible.
  3. It is not possible, for example, to dissolve sand in water, so this is not a spontaneous process. Water cannot flow uphill, so this is not a spontaneous process either. Some nonspontaneous processes can, however, be made to occur by providing energy from an external source.
  4. For instance, water can be made to flow uphill by using a pump. Similarly, though the electrolysis of water is not a spontaneous process, it can be made to occur by supplying electrical energy.

You may wonder about the difference between a nonspontaneous process which can be made to occur by supplying energy and a spontaneous process which requires initiation. The combination of gaseous hydrogen and oxygen to form water is a spontaneous reaction which has to be initiated by an electric spark.

∴ \(\mathrm{H}_2(\mathrm{~g})+\frac{1}{2} \mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{H}_2 \mathrm{O}(\mathrm{l})\)

  • The electrolysis of water is a nonspontaneous process which requires electrical energy. The difference is that once a spontaneous reaction has been initiated, it proceeds on its own, while a nonspontaneous process stops as soon as the supply of external energy stops.
  • Once the electric spark starts the reaction between hydrogen and oxygen, the process continues on its own. But if the supply of electric current is stopped during the electrolysis of water, the process stops.

Criterion for spontaneity: What makes certain processes possible or spontaneous, and certain others not possible or nonspontaneous? What is the driving force behind the occurrence of different processes? If we know this we can predict whether a certain process would be possible.

  • You have already read in several contexts that every system seeks to have the minimum possible energy in order to acquire the maximum stability. Water flows down a hill to have the minimum possible potential energy. A wound spring unwinds itself to minimise energy. Atoms form bonds with each other in order to have less energy.
  • This may lead you to the conclusion that the criterion for a process is the reduction of energy, or that a process is possible if it causes a reduction in the energy of the system. Let us consider some chemical reactions and see if this is right. In exothermic reactions, the energy of the products is less than the energy of the reactants and the reaction is accompanied by the evolution of heat.

⇒ \(\mathrm{C}(\mathrm{s})+\mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{CO}_2(\mathrm{~g}) \Delta H =-394 \mathrm{~kJ}\)

⇒ \(\mathrm{~N}_2(\mathrm{~g})+3 \mathrm{H}_2(\mathrm{~g}) \longrightarrow 2 \mathrm{NH}_3(\mathrm{~g}) \Delta H =-92 \mathrm{~kJ}\)

In such cases at least we may be right in saying that the tendency to attain a state of minimum energy (i.e., negative enthalpy change) is the driving force behind the occurrence of a reaction. But what about endothermic reactions which are spontaneous?

⇒ \(\mathrm{H}_2 \mathrm{O}(\mathrm{s}) \longrightarrow \mathrm{H}_2 \mathrm{O}(\mathrm{l}) \Delta H=5.86 \mathrm{~kJ}\)

⇒ \(\mathrm{NH}_4 \mathrm{Cl}(\mathrm{s})+\mathrm{aq} \longrightarrow \mathrm{NH}_4^{+}(\mathrm{aq})+\mathrm{Cl}^{-}(\mathrm{aq}) \Delta H=177.8 \mathrm{~kJ}\)

⇒ \(\mathrm{CaCO}_3(\mathrm{~s}) \longrightarrow \mathrm{CaO}(\mathrm{s})+\mathrm{CO}_3(\mathrm{~g}) \Delta H=15.1 \mathrm{~kJ}\)

  • These three reactions involve the absorption of heat, which means that the reactions lead to an increase in the energy of the system, so reduction of energy cannot be the only criterion determining the spontaneity of a reaction.
  • Reversible reactions also prove that decrease in energy or negative enthalpy change cannot be the only determining factor for the feasibility of a reaction. Suppose the forward reaction is exothermic then the backward reaction has to be endothermic and it would not occur if the only criterion for spontaneity were reduction of energy.

The tendency towards maximum disorder: To investigate what other factors could possibly determine the spontaneity of a process, let us consider a process for which ΔH = 0, since for such a process the energy factor could not be the driving force. The mixing of air and bromine vapour is just such a process.

  • Suppose vessel A contains bromine vapour and vessel B contains air and there is a movable partition between the two. Now suppose that the partition is removed. The two gases will mix completely. What makes this possible? When the samples of air and bromine vapour are in two separate vessels, there is more order and when they diffuse into each other, there is more disorder.
  • This tendency of a system to move towards greater disorder or to maximise randomness is the other driving force which makes reactions possible. It explains why endothermic reactions are possible despite the fact that ΔH is positive for such reactions. Take the case of the dissolution of ammonium chloride in water.

⇒ \(\mathrm{NH}_4 \mathrm{Cl}(\mathrm{s})+\mathrm{aq} \longrightarrow \mathrm{NH}_4^{+}(\mathrm{aq})+\mathrm{Cl}^{-}(\mathrm{aq})\)

  • This is an endothermic process, but it is possible because the randomness of the system increases. How? When the ions are held together in the crystal lattice there is more order and when they go into the aqueous solution, they are more free to move about, hence there is more disorder.
  • Just as the energy factor (minimising energy) cannot be the only criterion which determines the spontaneity of a process, nor can maximising randomness be the sole determining factor. If this were the case, the liquefaction of a gas (which increases order or decreases randomness) would not have been possible. But we will come to that later.

Basic Chemistry Class 11 Chapter 6 Thermodynamics Vessele A Contains Bromine Vapour And Vessele B Contains Air And Parition between Vesseles Is Removed Bromine And Air Mix

Entropy: The randomness or disorder of a system is measured in terms of a function called entropy. It is a state function, represented by S and measured in J K-1.

Change in entropy (ΔS) = Sfinal state – Sintial state

or ∑ Sproducts – ∑ Sreatants

When ΔS for a process is positive, the process leads to an increase in randomness or disorder.

In the examples discussed earlier, the processes are accompanied by an increase in entropy. We may again tend to conclude that for spontaneous processes there must be an increase in entropy. However, there are cases where entropy decreases. For example, the spontaneous condensation of steam to liquid water.

⇒ \(\underset{s^{\ominus}=188.7 \mathrm{JK}^{-1} \mathrm{~mol}^{-1}}{\mathrm{H}_2 \mathrm{O}(\mathrm{g})} \longrightarrow \underset{s^{\ominus}=70.0 \mathrm{JK}^{-1} \mathrm{~mol}^{-1}}{\mathrm{H}_2 \mathrm{O}(\mathrm{l})}\)

  • The standard molar entropy values for liquid water and water vapour are given and show that there is a decrease in entropy. Standard entropy may be defined in a manner similar to standard enthalpy.
  • The standard molar entropy of a substance is the entropy of one mole of the pure substance at 1 bar pressure and a specified temperature, usually 298 K. Let us not be in a hurry to make conclusions. When steam condenses, energy is given out, \(\Delta H^\ominus\) being -44.1 kJ mol-1 for the reaction.
  • Where does this energy go? If the water condenses on a window, the energy is passed on to the window and when it condenses in the air, energy is taken up by the air molecules. Therefore, the entropy of the surroundings increases.
  • The tendency towards disorder is summed up in the second law of thermodynamics, which states that spontaneous reactions are accompanied by a net increase in the entropy of the universe. Here the universe means the system and surroundings taken together,

⇒ \(\Delta S_{\text {univ }}>0 \) (2nd law thermodynamics).

⇒ \(\Delta S_{\text {uriv }}=\Delta S_{\text {sys }}+\Delta S_{\text {surr }}\)

In the case of condensation of water, \(\Delta S_{\text {univ }}=\Delta S_{\text {water }}+\Delta S_{\text {surr }}\)

  • The positive ΔSsurr outweighs the negative ΔSwater and the total entropy increases.
  • For all spontaneous processes, a stage is reached when the system attains equilibrium. A glass of hot water cools till it attains room temperature. It does not cool below room temperature. At this point, it is said to have attained equilibrium or, to be more precise, thermal equilibrium.

A gas expands till it fills the whole of the available space uniformly. Then no further expansion occurs. Again this is a state of equilibrium. At the point of equilibrium,

ΔSuniv = 0 (at equilibrium).

Having understood that entropy is a measure of disorder or randomness of a system, it may be easy to imagine that at absolute zero, the entropy of a perfectly crystalline substance would be zero. At any temperature above this, the entropy is positive and it increases with an increase in temperature.

For any substance Ssolid < Sliquid < Sgas

For any system the heat changes are reflected in the entropy of that system. Hence, let us find an expression relating entropy and heat. Consider an exothermic reaction. The heat released in an exothermic reaction flows to the surroundings and the heat lost by the system is equal to the heat gained by the surroundings.

qsurr = -qsys

This equation is valid even for an endothermic reaction. In the case of an exothermic reaction, the flow of heat into the surroundings increases its entropy, i.e.,

ΔSsurr>0.

The actual magnitude of ΔSsurr however, depends on the amount of heat released as well as the temperature. In fact, the effect of q on entropy is related to temperature. The addition of heat to colder surroundings has a greater effect on the entropy as compared to the effect produced on addition of heat to hotter surroundings. The effect of q is greater if the temperature is lower. This can be mathematically expressed as

⇒ \(\Delta S_{\text {surt }}=\frac{q_{\text {surr }}}{T}=-\frac{q_{\text {sys }}}{T} .\)….(1)

We know that at equilibrium \(\Delta S_{\text {undv }}=0=\Delta S_{\text {sys }}+\Delta S_{\text {surr }} \text {. }\)….. (2)

Therefore, form equations (1) and (2) we get \(\Delta S_{s y s}=\frac{+q_{s y s}}{T}\)

If the process is reversible, the change in entropy of the system is given by \(\Delta S_{r v}=\frac{q_{r v v}}{T}.\)

  • You may wonder that if no real processes take place reversibly, what is the use of the above equation. Although transfer of heat between system and surroundings is almost impossible to achieve in a reversible manner, this idealised path is important for the definition of ΔS.
  • Entropy is a state function and the value of ΔS is the same for a change from say any state A to any other state B, irrespective of the path taken. Whether the change is carried out in a reversible or an irreversible manner, ΔS is the same.
  • Thus whatever may be the actual path taken, ΔS is calculated using qrev only. qrev is the heat associated with the process had it traversed a reversible path.

The overall criterion for spontaneity: To get back to what we were discussing earlier, the overall criterion for the spontaneity of a process is the resultant of two tendencies—

  1. The tendency towards minimum energy and
  2. The tendency towards maximum entropy. These two tendencies act independently of each other and may work in the same or opposite directions.

The resultant of the two tendencies is expressed by another thermodynamic function called Gibbs free energy (G). The free energy of a system is a measure of its capacity to do useful work. The enthalpy (H), entropy (S), and free energy (G) of a system are related by the following expression.

G = H-TS (here T is the absolute temperature).

At constant temperature and pressure, the change in free energy (AG) during a process can be obtained by the following equation.

ΔG = ΔH -TΔS…… Equation 7

  1. The change in free energy takes into account both the change in enthalpy and the change in entropy, and is the factor which determines the spontaneity of a process. You know that a process is possible if it is accompanied by a decrease in energy (ΔH negative) and an increase in entropy (ΔS positive).
  2. Some processes for which ΔH is positive are possible because they are accompanied by an increase in entropy (ΔS positive). Some processes for which ΔS is negative are possible because they are accompanied by a decrease in energy (ΔH negative).
  3. Obviously, the conditions that are most favourable for the occurrence of a process are a decrease in energy and an increase in entropy, i.e., when ΔH is negative and ΔS is positive.

This makes ΔG negative. Basically, the criterion for a spontaneous reaction or feasibility of a reaction is that ΔG should be negative. A spontaneous reaction is always accompanied by a decrease in free energy, i.e.,

ΔG < 0 for a spontaneous process.

The tendency of the system to undergo a change in the direction of decreasing free energy makes it reach a state when ΔG = 0 at some point. This is the equilibrium state.

Thus, ΔG > 0 for a nonspontaneous process.

  • However, the reverse of a nonspontaneous process (for which ΔG is positive) is spontaneous.
  • In order to understand the meaning of free energy for a process being zero, let us consider the simple case of water. Water freezes spontaneously at T<0°C and icc melts spontaneously at T>0°C. At T = 0°C and 1 atm pressure, the ΔG for melting of ice and freezing of water is zero.
  • At this temperature, theoretically, ice should not melt and water should not freeze. This means the two phases, liquid and solid, may coexist in equilibrium with each other.
  • We may also have chemical reactions for which ΔG = 0. As a spontaneous chemical reaction proceeds (ΔG < 0), reactants turn into products and ΔG becomes less negative. At a point ΔG becomes zero—no more products are formed nor are any reactants consumed. This is a state of equilibrium or in fact chemical equilibrium.
  • Reactions with negative ΔG are called exergonic and those with positive ΔG are called endergonic reactions.

When is ΔG negative? It is worth finding an answer to this question because it will help us determine which processes are feasible and under what conditions. The figure represents graphically the relation between ΔH, ΔS, and ΔG.

  1. Obviously, when ΔH is negative and ΔS is positive, ΔG will be negative (line ‘a’)
  2. When ΔH is negative and ΔS, too, is negative, ΔG will be negative if the magnitude of ΔH is greater than that of TΔS (line ‘b’), i.e., below temperature Tb.
  3. When both ΔH and ΔS are positive, AG will be negative if the magnitude of TΔS is greater than that of ΔH (line ‘c’, i.e., above temperature Tc).
  4. When ΔH is positive, while ΔS is negative, ΔG will be positive and the reaction will not be spontaneous.

Effect of temperature on spontaneity If you think about the last two conditions for ΔG to be negative, it will strike you that temperature must be an important factor in determining the spontaneity of a reaction. Say, both ΔH and ΔS are negative for a process. If tire magnitude of ΔH is greater than that of TΔS, ΔG will be negative and the process will be feasible.

  • It could be possible that this is true up to a certain value of T, beyond which TΔS becomes greater in magnitude than ΔH. Similarly, if both ΔH and ΔS are positive for a particular process, there could be a temperature below which the magnitude of TΔS is less than that of ΔH, and the process is not feasible.
  • For an endothermic process, ΔH is positive, i.e., the energy factor opposes the process. For such a process to occur, TΔS must be positive and its magnitude must be greater than that of ΔH.

Basic Chemistry Class 11 Chapter 6 Thermodynamics Plots Of Abgle G Vas Temperature Under Various Conditions

  • For such a process, a higher temperature is more favorable as tills increases the magnitude of TΔS. In fact, there may be a temperature below which the magnitude of TΔS becomes less than that of ΔH and the process becomes unfeasible.
  • For an exothermic process, ΔH is negative and ΔG is negative at all temperatures if ΔS is positive. However, if ΔS is negative, the process is feasible only if the magnitude of TΔS is less than that of ΔH.
  • Obviously, such a reaction would be more feasible at a low temperature and may even become nonspontaneous above a particular temperature.

If ΔH is the major factor contributing to negative ΔG value, the reaction is called ‘enthalpy driven’ and if entropy is the factor behind it, it is called an ‘entropy driven’ reaction. A few values of \(\Delta H^{\ominus}, \Delta G^{\ominus} \text { and } \Delta S^\ominus\) are given in Table. Let us now summarise whatever we have discussed so far pertaining to the feasibility of a reaction in Table.

Units of \(\Delta G^{\ominus}\): cal mol-1 in the CGS system and J mol-1 or kJ mol-1 in the SI system.

Basic Chemistry Class 11 Chapter 6 Thermodynamics Standard Enthalpies, Entrapies And Free Energies Of Some Common Reactions At 298 K

Basic Chemistry Class 11 Chapter 6 Thermodynamics Relation Between Spontaneity And Signs OF Delta H And Delts S And Delta G

Standard free energy change: The standard free energy change of a reaction is defined as the free energy change at a specified temperature when the reactants in their standard states are converted to products in their standard states. It is denoted by \(\Delta G^{\ominus}\).

Calculation of \(\Delta G^{\ominus}\)° of a reaction Free energy is a state function. Hence, similar to \(\Delta H^{\ominus}\), \(\Delta G^{\ominus}\) for a reaction can be calculated from the free energy of formation of reactants and products.

The standard free energy of formation of an element in its standard state is taken as zero. Table lists values of standard enthalpy of formation, the standard free energy of formation, and absolute standard entropies of some compounds.

Tire standard free energy change of a reaction is obtained by subtracting the sum of the standard free energy of formation of the reactants from that of the products.

∴ \(\Delta_r G^{\ominus}=\Sigma v \Delta_f G^{\ominus} \text { (products) }-\Sigma v \Delta_f G^{\ominus} \text { (reactants) }\)

where v’s are the stoichiometric coefficients of the products and reactants respectively.

∴ \(\Delta G^{\ominus}\) may also be calculated from \(\Delta H^{\ominus}\) and \(\Delta S^{\ominus}\) of the reaction using an equation similar to Equation 7.

∴ \(\Delta_f G^{\ominus}\) = \(\Delta_f G^{\ominus}\) at a particular temperature.

We have seen how to calculate \(\Delta G^{\ominus}\) for a reaction. However, all chemical reactions do not occur with the reactants and products in their standard states. We can determine AG for a reaction of this kind using \(\Delta G^{\ominus}\) of the same reaction at standard conditions and the reaction quotient of the reaction. Consider the reaction

αA + βB → γC + δD

A, B, C, and D are the reactants and products while α,β,γ, and δ represent the respective stoichiometries.

The reaction quotient, Qc, of this reaction is given by \(Q_c=\frac{[C]^\gamma[D]^8}{[A]^\alpha[B]^\beta}\)

The square brackets indicate the concentrations of the reactant/product. Note that the products appear in the numerator and reactants appear in the denominator. The concentration terms of the species are raised to their respective stoichiometric coefficients.

ΔG at any temperature T is related to \(\Delta G^{\ominus}\) as follows:

ΔG = \(\Delta G^{\ominus}\) + RT In Qc ……. Equation 8

where R is the gas constant.

If the species involved in the reaction are gaseous in nature, concentration is replaced by partial pressure, and the reaction quotient is represented as Qp.

At equilibrium, Q = K, the equilibrium constant. (Qc = Kc or Qp = Kp, as the case may be.)

Basic Chemistry Class 11 Chapter 6 Thermodynamics Standard Enthalpies Of Formation Standard Free Energies Of Formation And Absloute Standard Entropies

The Equilibrium constant K of the reaction αA + βB → γC + δD  is given by

K = \(\frac{[\mathrm{C}]_e^r[\mathrm{D}]_e^5}{[\mathrm{~A}]_e^\alpha[\mathrm{B}]_e^\beta}\)

where the subscripts e denote equilibrium concentrations. Also, at equilibrium, ΔG = 0.

Hence, \(\Delta G^{\ominus}\) = -RT In K = -2.303 RT log K.

This gives yet another method of determining \(\Delta G^{\ominus}\) of a reaction. The pressure should be expressed in bar and the concentration in mol L-1 while determining Kp and Kc respectively.

Example 1. Calculate the entropy change AS per mole for the following reactions.

1. Combustion of hydrogen at 298 K \(\mathrm{H}_2(\mathrm{~g})+\frac{1}{2} \mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{H}_2 \mathrm{O}(\mathrm{g})\)

\(\Delta H=-241.6 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

\(\Delta G=-228.4 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

2. Vaporisation of methanol at its normal boiling point

Methanol (l) → Methanol (g), Δvap H = 23.9 kJ mol-1, Boiling point = 338 K
Solution:

1. ΔG = ΔH – TΔS.

∴ \(\Delta S=\frac{\Delta H-\Delta G}{T}=\frac{-2416-(-228.4)}{298}=-0.044 \mathrm{~kJ} \mathrm{~K}^{-1} \mathrm{~mol}^{-1} .\)

2. ΔG = ΔH -TΔS.

ΔG = 0.

∴ \(\Delta S=\frac{\Delta_{\mathrm{vap}} H}{T}=\frac{23.9}{338}=0.071 \mathrm{~kJ} \mathrm{~K}^{-1} \mathrm{~mol}^{-1} \text {. }\)

Example 2. Calculate the free energy change per mole for the following reactions and predict the feasibility of the reactions.

1. \(\mathrm{CaCO}_3(\mathrm{~s}) \longrightarrow \mathrm{CaO}(\mathrm{s})+\mathrm{CO}_2(\mathrm{~g}) \quad \text { at } 298 \mathrm{~K}\)

⇒ \(\Delta H=177.9 \mathrm{~kJ} \mathrm{~mol}^{-1} \quad \Delta S=160.4 \mathrm{JK}^{-1} \mathrm{~mol}^{-1}\)

2. \(2 \mathrm{NO}_2(\mathrm{~g}) \longrightarrow \mathrm{N}_2 \mathrm{O}_4(\mathrm{~g}) \quad \text { at } 298 \mathrm{~K}\)

⇒ \(\Delta \mathrm{S}=175.6 \mathrm{JK}^{-1} \mathrm{~mol}^{-1} \quad \Delta H=-57.2 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

Solution:

1. ΔG =ΔH – TΔS.

∴ ΔG = 177.9 x 103 – 298 x 160.4 = 130100.8 J mol-1 = 130.1 kJ mol-1.

ΔG is positive. Therefore, the reaction is nonspontaneous at this temperature.

2. ΔG = -57.2×103 – (298 x 175.6) =-109528.8 J mol-1 =-109.5 kj mol”1.

ΔG is negative. Therefore, this reaction is spontaneous at 298 K.

Why Should There Be An Energy Crisis

While reading about the law of conservation of energy, it may have struck you as strange that there should be an energy crisis in the world when energy is not really lost in any process, it is merely converted from one form into another. It is true that energy is merely converted from one form into another, but it is useful to us only in certain forms.

  • In most processes that we encounter, a portion of energy is converted into forms we cannot use. Let us look at it in another way. Most of our energy requirements are met by the burning of fossil fuels. The combustion of fuels produces energy which we use. But it also produces carbon dioxide and water, which escape into the air.
  • The combustion of fuels is a spontaneous process, but the reverse is not a spontaneous process. So, the combustion of fuels is a one-sided process, which increases disorder. It is this tendency of most processes to increase disorder that wastes energy in our terms.
  • Even the energy released during the combustion of fuels cannot be utilised fully to perform useful work. A part of it goes to increase disorder or is wasted according to us. This is why the efficiency of an engine is always less than 1.

∴ \(\frac{\text { work done by engine }}{\text { energy fed into engine }}<1\)

This follows from the equation

ΔG = ΔH-TΔS

or ΔH = ΔG + TΔS.

  • If ΔH is the total energy fed into an engine then some of it, i.e., TΔS, will be used to increase disorder and only a part of it (ΔG) will be available for useful work. This is why, though the total energy of the world remains constant, every living process converts a part of it into a form that cannot be used.
  • This is the part that goes to increase disorder in the universe. Since AG is the useful part of energy that can be harnessed, it is a measure of the maximum amount of energy that is ‘free’ to do useful work. Hence its name.

Example 1. Calculate the standard enthalpy of formation of ethyl alcohol from the following data.

  1. \(\mathrm{C}_2 \mathrm{H}_5 \mathrm{OH}(\mathrm{l})+3 \mathrm{O}_2(\mathrm{~g}) \longrightarrow 2 \mathrm{CO}_2(\mathrm{~g})+3 \mathrm{H}_2 \mathrm{O}(\mathrm{l}) \Delta_{\mathrm{f}} \mathrm{H}_1^{\ominus}=-1368 \mathrm{~kJ} \mathrm{~mol}^{-1}\) ….. (1)
  2. \(\mathrm{C}(\mathrm{s})+\mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{CO}_2(\mathrm{~g}) \Delta_{\mathrm{f}} H_2^{\ominus}=-393.5 \mathrm{~kJ} \mathrm{~mol}^{-1}\) …..(2)
  3. \(\mathrm{H}_2(\mathrm{~g})+\mathrm{S}_2 \mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{H}_2 \mathrm{O}(\mathrm{l}) \Delta_{\mathrm{f}} H_3^{\ominus}=-286.0 \mathrm{~kJ} \mathrm{~mol}^{-1}\)…..(3)

Solution: The required equation is

∴ \(2 \mathrm{C}(\mathrm{s})+3 \mathrm{H}_2(\mathrm{~g})+\frac{1}{2} \mathrm{O}_2 \rightarrow \mathrm{C}_2 \mathrm{H}_5 \mathrm{OH}(\mathrm{l}) \quad \Delta_{\mathrm{f}} \mathrm{H}^{\ominus}=\)

To obtain this equation, first multiply Equation (2) by 2 and Equation (3) by 3 and add the products to get Equation (4). Then subtract Equation (1) from Equation (4),

Basic Chemistry Class 11 Chapter 6 Thermodynamics Standard Enthalpy Of Formatiom Of Ethyl Alcohol

∴ \(\Delta_{\mathrm{f}} H^{\ominus}=\left[2 \times \Delta_f H_2^{\ominus}+3 \times \Delta_{\mathrm{f}} H_3^{\ominus}\right]-\Delta_{\mathrm{f}} H_1^{\ominus}\)

= [(2 x – 393.5) + (3 x – 286.0)] – [-1368] = -1645 +1368 = -277 kJ mol-1.

Example 2. Calculate the heat of formation of methane, given that the heats of combustion of methane, graphite and hydrogen are – 890, – 394, – 286 kJ mol-1 respectively.
Solution:

The required equation is \(\mathrm{C}(\mathrm{s})+2 \mathrm{H}_2(\mathrm{~g}) \longrightarrow \mathrm{CH}_4(\mathrm{~g}) \quad \Delta_{\mathrm{f}} \mathrm{H}=\text { ? }\)

The equation for the combustion of methane is \(\mathrm{CH}_4(\mathrm{~g})+2 \mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{CO}_2(\mathrm{~g})+2 \mathrm{H}_2 \mathrm{O}(\mathrm{g}) \quad \Delta_{\mathrm{c}} \mathrm{H}_1=-890 \mathrm{~kJ} \mathrm{~mol}^{-1}\) ….(1)

The equation for the combustion of graphite is \(\mathrm{C}(\mathrm{s})+\mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{CO}_2(\mathrm{~g}) \quad \Delta_{\mathrm{c}} \mathrm{H}_2=-394 \mathrm{~kJ} \mathrm{~mol}^{-1}\) …..(2)

The equation for the combustion of hydrogen is \(\mathrm{H}_2(\mathrm{~g})+\frac{1}{2} \mathrm{O}_2 \longrightarrow \mathrm{H}_2 \mathrm{O}(\mathrm{g}) \quad \Delta_{\mathrm{c}} \mathrm{H}_3=-286 \mathrm{~kJ} \mathrm{~mol}^{-1}\)…. (3)

Multiply Equation (3) by 2 and add the product to Equation (2).

Basic Chemistry Class 11 Chapter 6 Thermodynamics Heat Formatiom Of Methane

This is the required equation.

Then ΔfH =[ΔcH2 + (2 X ΔcH3)]-[ΔcH1]

= ((-394) + (2 x – 286)] – [-890] = -394-572 + 890 = -76 kJ mol-1.

Example 3. Calculate the heat of reaction of \(\mathrm{CO}_2(\mathrm{~g})+\mathrm{H}_2(\mathrm{~g}) \longrightarrow \mathrm{CO}(\mathrm{g})+\mathrm{H}_2 \mathrm{O}(\mathrm{l})\)

Given that

  1. \(\mathrm{C}(\mathrm{s})+\frac{1}{2} \mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{CO}(\mathrm{g}) \Delta_f H_2=-110.35 \mathrm{~kJ}\)
  2. \(\mathrm{C}(\mathrm{s})+\mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{CO}_2(\mathrm{~g}) \Delta_f H_2=-393.35 \mathrm{~kJ}\)
  3. \(\mathrm{H}_2(\mathrm{~g})+\frac{1}{2} \mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{H}_2 \mathrm{O}(\mathrm{l}) \Delta_f H_3=-241.6 \mathrm{~kJ}\)

Solution:

This problem can be solved by two methods.

1. Applying Hess’s law:

Add Equation (1) and Equation (3) and then subtract Equation (2) from the resultant equation.

⇒ \(\begin{aligned}
\mathrm{C}(\mathrm{s})+\mathrm{H}_2(\mathrm{~g})+\mathrm{O}_2(\mathrm{~g}) & \longrightarrow \mathrm{CO}(\mathrm{g})+\mathrm{H}_2 \mathrm{O}(\mathrm{l}) \\
-\mathrm{C}\left(\mathrm{C}(\mathrm{s})+\mathrm{O}_2(\mathrm{~g})\right. & \left.\longrightarrow \mathrm{CO}_2(\mathrm{~g})\right) \\
\hline \mathrm{H}_2(\mathrm{~g})+\mathrm{CO}_2(\mathrm{~g}) \longrightarrow & \longrightarrow \mathrm{CO}(\mathrm{g})+\mathrm{H}_2 \mathrm{O}(\mathrm{l})
\end{aligned}\)

This is the required equation.

∴ ΔrH = ΔfH1 – ΔfH3 – ΔfH2 =(-110.35) + (- 2416)- (-39335) = + 414 kj.

2. ΔfH = ∑ΔfH(products) – ∑ΔfH(reactants)

= [ΔfH(CO) + ΔfH(H2O)]- ΔfH(CO2)

= -110.35 + (- 2416)- (-39335) = + 414 kj.

Example 4. Just before an athletic event, a participant is given 100 g of glucose (C6H12O6), whose energy equivalent is 1560 kj. He utilises 50% of this gained energy in the event. Calculate the weight of water he would need to perspire in order to avoid storing extra energy in the body, ifthe enthalpy ofevaporation ofwater is 44 kj mol-1.
Solution:

Energy from 100 g of glucose = 1560 kJ.

Energy utilised = 50% of1560 = 780 kJ.

Energy not used = 1560- 780 = 780 kJ.

Enthalpy of evaporation of water = 44 kJ mol-1.

44 kJ of energy evaporates1 mol of water.

780 kJ of energy would evaporate \(\frac{1 \times 780}{44}\) =17.73 mol of water.

17.73 mol of water = 17.73 x18 g = 319.14 g of water.

Example 5. Calculate the bond energy of the Cl—Cl bondfrom the equation \(\mathrm{CH}_4(\mathrm{~g})+\mathrm{Cl}_2(\mathrm{~g}) \longrightarrow \mathrm{CH}_3 \mathrm{Cl}(\mathrm{g})+\mathrm{HCl}(\mathrm{g}) \Delta \mathrm{H}=-100.3 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

given that the bond energies ofC—H, C—Cl, H—Cl bonds are 413, 326 and 431 kJ mol-1 respectively.

Solution: The given equation can be written as

Basic Chemistry Class 11 Chapter 6 Thermodynamics Bond Energy C-H, C-Cl, H-Cl Bonds

∴ \(\Delta H=\left[4 \Delta H_{\mathrm{C}-\mathrm{H}}+\Delta H_{\mathrm{C}-\mathrm{a}}\right]-\left[3 \Delta H_{\mathrm{C}-\mathrm{H}}+\Delta H_{\mathrm{C}-\mathrm{a}}+\Delta H_{\mathrm{H}-\mathrm{a}}\right]\)

or \(-100.3=\left[(4 \times 413)+\Delta H_{\mathrm{a}-\mathrm{a}}\right]-[(3 \times 413)+326+431]\)

∴ \(\quad \Delta H_{\mathrm{a}-\mathrm{C}}=243.7 \mathrm{~kJ} \mathrm{~mol}^{-1}\).

Example 6. Calculate the enthalpy change for the reaction \(\mathrm{H}_2(\mathrm{~g})+\mathrm{Br}_2(\mathrm{~g}) \longrightarrow 2 \mathrm{HBr}(\mathrm{g})\) given that the bond enthalpies of H—H, Br—Br and H—Br are 435,192 and 364 kj mol-1 respectively.
Solution:

ΔH =∑ bond enthalpies of reactants – ∑ bond enthalpies of products

= [ΔHH-H + ΔHBr-Br] – [2 x ΔHH-Br]

= 435 + 192-(364 x 2)

= 627-728 =-101 kj mol-1.

Example 7. Calculate the Gibbs energy change for the reaction \(4 \mathrm{Fe}(\mathrm{s})+3 \mathrm{O}_2(\mathrm{~g}) \rightleftharpoons 2 \mathrm{Fe}_2 \mathrm{O}_3(\mathrm{~s})\) at 25°C when O2 is at a partial pressure of 1.5 bar. \(\Delta_r G^e\) =-1484.4 kJ mol-1

Solution:

The reaction quotient Q = \(\frac{\left[\mathrm{Fe}_2 \mathrm{O}_3(\mathrm{~s})\right]^2}{[\mathrm{Fe}(\mathrm{s})]^4\left[p_{\mathrm{O}_2}\right]^3}\).

Since the concentrations of pure solids and liquids are constant,

Q = \(\frac{1}{\left(p_{\mathrm{O}_2}\right)^3}=\frac{1}{(1.5)^3}=0.2963\)

Now, \(\Delta_r G=\Delta_r G^\ominus\) + RT ln Q

= -1484.4 x 103 + 8.314 x 298 x ln (0.2963) (R = 8.314 JK-1 mol-1)

= – 1,484,400-3013.7 =-1,487,413.7 J mol-1 = -1487.4 kJ mol-1.

Example 8. Arrange the following systems in order of increasing entropy (compare one mole of each).

1. \(\mathrm{H}_2 \mathrm{O}$ (l), \mathrm{H}_2 \mathrm{O} (s), \mathrm{H}_2 \mathrm{O} (g)\)

2. \(\mathrm{H}_2(\mathrm{~g}), \mathrm{HBrO}_4(\mathrm{~g}), \mathrm{HBr}(\mathrm{g})\)

Solution:

Entropy varies as

1. \(\mathrm{H}_2 \mathrm{O}(\mathrm{s})<\mathrm{H}_2 \mathrm{O}(\mathrm{l})<\mathrm{H}_2 \mathrm{O}(\mathrm{g})\)

2. \(\mathrm{H}_2(\mathrm{~g})<\mathrm{HBr}(\mathrm{g})<\mathrm{HBrO}_4(\mathrm{~g})\)

If the physical state is the same, the more complicated the molecule, the more is the entropy.

Example 9. For the reaction \(2 \mathrm{SO}_2(\mathrm{~g})+\mathrm{O}_2(\mathrm{~g}) \longrightarrow 2 \mathrm{SO}_3(\mathrm{~g})\) find ΔG at 25°C. The partial pressures of SO2, O2 and SO3 are 1.0 bar, 0.5 bar and 0.1 bar respectively. \(\Delta G^{\ominus}\) = -141.8kj mol-1.
Solution:

⇒ \(\Delta G=\Delta G^{\ominus}\) + RT In Qf; where Qf, is the reaction quotient.

⇒ \(Q_p=\frac{\left(p_{\mathrm{SO}_3}\right)^2}{\left(p_{\mathrm{SO}_2}\right)^2\left(p_{\mathrm{O}_2}\right)}=\frac{(0.1)^2}{(1)^2(0.5)}=0.02\)

∴ \(\Delta G=\Delta G^{\ominus}+R T \ln Q\)

= -1418 x 103 + 8314 x 298 x In 0.02

=-151,4923 J mol-1

=-151.5 kj mol-1

Example 10. The \(\Delta S_{\text {system }} \text { and } \Delta_t H^{\ominus}\) for the following reaction at 25 C are 220.5 J k-1 mol-1 and 90.7 kJ mol-1. Find S and say in which direction the reaction is not spontaneous.

\(\mathrm{CH}_3 \mathrm{OH}(\mathrm{g}) \rightleftharpoons \mathrm{CO}(\mathrm{g})+2 \mathrm{H}_2(\mathrm{~g})\)

Solution:

⇒ \(\Delta S_{\text {total }}=\Delta S_{\text {sys }}+\Delta S_{\text {surr }} \text {. }\)

⇒ \(\Delta S_{\text {sys }}=220.8 \mathrm{~J} \mathrm{~K}^{-1} \mathrm{~mol}^{-1} \text {. }\)

⇒ \(\Delta S_{\text {surr }}=\frac{-\Delta H^{\ominus}}{T}\)

= \(-\frac{\left(90.7 \times 10^3\right)}{298} \mathrm{~J} \mathrm{~K}^{-1} \mathrm{~mol}^{-1}\)

= \(-304.4 \mathrm{~J} \mathrm{~K}^{-1} \mathrm{~mol}^{-1} \text {. }\)

\(\Delta S_{\text {total }}=220.8-304.4=-83.6 \mathrm{~J} \mathrm{~K}^{-1} \mathrm{~mol}^{-1}\)

As the entropy decreases (ΔS = -ve) the reaction is not spontaneous in the forward direction but the reverse reaction is spontaneous.

Example 11. Considering only entropy,find whether thefollowing reaction is spontaneous or not at 25°C

⇒ \(\mathrm{N}_2(\mathrm{~g})+2 \mathrm{H}_2(\mathrm{~g}) \rightleftharpoons \mathrm{N}_2 \mathrm{H}_4(\mathrm{l})\)

⇒ \(\mathrm{N}_2(\mathrm{~g}) \quad \mathrm{H}_2(\mathrm{~g}) \quad \mathrm{N}_2 \mathrm{H}_4(\mathrm{l})\)

Given \(\begin{array}{lccc}
\Delta H^{\ominus}\left(\mathrm{kJ} \mathrm{mol}^{-1}\right) & 0 & 0 & 50.6 \\
\Delta S^{\ominus}\left(\mathrm{J} \mathrm{K}^{-1} \mathrm{~mol}^{-1}\right) & 191.5 & 130.6 & 121.2
\end{array}\)

Solution: To predict whether or not a reaction is spontaneous, we must find the sign of \(\Delta S_{\text {total }}=\Delta S_{\text {sys }}+\Delta S_{\text {surr }}\)

The entropy change of the system is equal to the standard entropy of the reaction and

⇒ \(\Delta S_{\text {surr }}=-\Delta_r H^{\ominus} / T \text {. }\)

⇒ \(\Delta S_{\mathrm{sys}}=S^{\ominus}\left(\mathrm{N}_2 \mathrm{H}_4(\mathrm{l})\right)-S^{\ominus}\left(\mathrm{N}_2(\mathrm{~g})\right)-2 S^{\ominus}\left(\mathrm{H}_2(\mathrm{~g})\right)\)

= \(121.2-191.5-(2 \times 130.6)\)

= \(-3315 \mathrm{~J} \mathrm{~K}^{-1} \mathrm{~mol}^{-1} \text {. }\)

⇒ \(\Delta_{\mathrm{r}} H^{\ominus}=\Delta_f H^{\ominus}\left(\mathrm{N}_2 \mathrm{H}_4(\mathrm{l})\right)-\Delta_{\mathrm{f}} H^{\ominus}\left(\mathrm{N}_2(\mathrm{~g})\right)-2 \Delta_{\mathrm{f}} H^{\ominus}\left(\mathrm{H}_2(\mathrm{~g})\right)\)

= \(50.6-0-0=50.6 \mathrm{~kJ}^{-1} \mathrm{~mol}^{-1} \text {. }\)

⇒ \(\Delta S_{\text {surr }}=-\frac{\Delta_r H^{\ominus}}{T}=\frac{-50.6}{25+273}=-0.1698 \mathrm{~kJ} \mathrm{~K}^{-1} \mathrm{~mol}^{-1}=-169.8 \mathrm{~J} \mathrm{~K}^{-1} \mathrm{~mol}^{-1} \text {. }\)

⇒ \(\Delta S_{\text {total }}=\Delta S_{\text {sys }}+\Delta S_{\text {surt }}\)

= \(-331.5+(-169.8)\)

= \(-5013 \mathrm{~J} \mathrm{~K}^{-1} \mathrm{~mol}^{-1} \text {. }\)

As \(\Delta S_{\text {total }}\) is negative, the reaction is not spontaneous.

Example 12. The values of equilibrium constant for the following reaction at two different temperatures are 168 x 10-5 and 0.0123 respectively. If the first temperature is 1000°C, find the second temperature.

\(\mathrm{H}_2 \mathrm{O}(\mathrm{g})+\mathrm{C}(\mathrm{s}) \longrightarrow \mathrm{H}_2(\mathrm{~g})+\mathrm{CO}(\mathrm{g})\)

Solution:

⇒ \(\Delta G^{\ominus}=-R T_1 \ln K_{p_1}\) …. (1)

⇒ \(\Delta G^{\ominus}=-R T_2 \ln K_{p_2}\)…. (2)

where T1 and T2 are the two different temperatures with the corresponding equilibrium constants \(K_{p_1}\) and \(K_{p_2}\) . Equating RHS of (1) and (2),

∴ \(R T_1 \ln K_{p_1}=R T_2 \ln K_{p_2}\)

or \(T_1 \ln K_{p_1}=T_2 \ln K_{p_2}\)

or \(T_2=T_1 \cdot \frac{\ln K_{p_1}}{\ln K_{p_2}}=(1273) \times \frac{\ln \left(168 \times 10^{-5}\right)}{\ln (0.0123)}=3182 \mathrm{~K}\) .

Example 13. Calculate Kp for thefollowing reactions at 298 K

  1. \(2 \mathrm{SO}_2(\mathrm{~g})+\mathrm{O}_2(\mathrm{~g}) \longrightarrow 2 \mathrm{SO}_2(\mathrm{~g}) \Delta G^{\ominus}=-142 \mathrm{~kJ} \mathrm{~mol}^{-1}\)
  2. \(\mathrm{SO}_3(\mathrm{~g})+\mathrm{H}_2 \mathrm{O}(\mathrm{l}) \longrightarrow \mathrm{H}_2 \mathrm{SO}_4(\mathrm{l}) \Delta G^{\ominus}=-81.7 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

Solution:

1. \(\Delta G^{\ominus}=-R^{\prime} T \ln K_r \) =\(-2.303 R T \log K_p\)

⇒ \(\log K_p=-\frac{\Delta G^{\ominus}}{2.303 R T}\)

⇒  \(\log K_p=-\frac{\Delta G^{\ominus}}{2.303 R T}\)

or \(K_p=\operatorname{Antilog}\left(\frac{-\Delta G^{\ominus}}{2.303 R T}\right)\)

= \(\text { Antilog }\left(\frac{142,000}{2.303 \times 8.314 \times 2.98}\right)=7.7 \times 10^{24}\).

2. \(K_p= Antilog \left(\frac{81,700}{2.303 \times 8.314 \times 298}\right)=2.08 \times 10^4\).

Example 14. Calculate \(\Delta G^{\ominus}\) for thefollowing reactions

1. \(\mathrm{CaCO},(\mathrm{s}) \longrightarrow \mathrm{CaO}(\mathrm{s})+\mathrm{CO}_2(\mathrm{~g})\)

⇒  \(K_p=1.06 \text { at } 500^{\circ} \mathrm{C}\)

2. \(\mathrm{H}_2(\mathrm{~g})+\mathrm{I}_2(\mathrm{~g}) \rightleftharpoons 2 \mathrm{HI}(\mathrm{g})\)

⇒  \(\mathrm{K}_p=98.9 \text { at } 300^{\circ} \mathrm{C}\).

3. \(\mathrm{CoO}(\mathrm{s})+\mathrm{CO}(\mathrm{g}) \rightleftharpoons \mathrm{Co}(\mathrm{s})+\mathrm{CO}_2(\mathrm{~g})\)

⇒  \(K_p=1.09 \times 10^4 \text { at } 277^{\circ} \mathrm{C}\) .

Solution: Since \(\Delta G^{\ominus}\), substituting the values of R, T and Kp we get \(\Delta G^{\ominus}\) for each reaction.

1. \(\Delta G^{\ominus}\) = -8.314 x (500 + 273) x ln(1.06) = -374.4 J mol-1.

2. \(\Delta G^{\ominus}\) = -8.314 x (300 + 273) x ln(98.9) = -21,885.9 J mol-1 = -21.88 kj mol-1.

3. \(\Delta G^{\ominus}\) = -8.314 x (277 + 273) x ln(1.09 x 104 )\(\Delta G^{\ominus}\)= -42310.2 J mol-1 = -42.51 kJ mol-1.

Example 15. Calculate \(\Delta G^{\ominus}\) for the following reactions which are carried out under standard state conditions.

  1. \(2 \mathrm{LiOH}(\mathrm{s})+\mathrm{CO}_2(\mathrm{~g}) \longrightarrow \mathrm{Li}_2 \mathrm{CO}_3(\mathrm{~s})+\mathrm{H}_2 \mathrm{O}(\mathrm{g})\)
  2. \(\mathrm{H}_2(\mathrm{~g})+\mathrm{Br}_2(\mathrm{I}) \longrightarrow 2 \mathrm{HBr}(\mathrm{g})\)
  3. \(\mathrm{CaO}(\mathrm{s})+\mathrm{H}_2 \mathrm{O}(\mathrm{l}) \rightarrow \mathrm{Ca}(\mathrm{OH})_2(\mathrm{~s})\)
  4. \(2 \mathrm{HgO}(\mathrm{s}, \text { red }) \rightarrow 2 \mathrm{Hg}(\mathrm{l})+\mathrm{O}_2(\mathrm{~g})\)

The standardfree energies of the compounds are given below.

⇒ \(\Delta_i G\left(\mathrm{~kJ} \mathrm{~mol}^{-1}\right)\)

⇒  \(\mathrm{LiOH}(\mathrm{s}) -443.9 \)

⇒  \(\mathrm{CO}_2(\mathrm{~g}) -394.4\)

⇒ \(\mathrm{Li}_2 \mathrm{CO}_3(\mathrm{~s}) -1132.4\)

⇒ \(\mathrm{H}_2 \mathrm{O}(\mathrm{g}) -228.6\)

⇒  \(\mathrm{H}_2(\mathrm{~g}) 0\)

⇒ \(\mathrm{Br}_2(\mathrm{l}) 0\)

⇒ \(\mathrm{HBr}(\mathrm{g}) -53.43\)

⇒ \(\mathrm{CaO}(\mathrm{s}) -604.2\)

⇒ \(\mathrm{H}_2 \mathrm{O}(\mathrm{l}) -237.2\)

⇒ \(\mathrm{Ca}(\mathrm{OH})_2(\mathrm{~s}) -896.8\)

⇒ \(\mathrm{HgO}(\mathrm{s}, \text { red) } -58.55\)

⇒ \(\mathrm{Hg}(\mathrm{l}) 0\)

⇒ \(\mathrm{O}_2(\mathrm{~g}) 0\)

⇒ \(\Delta_1 \mathrm{H}(\mathrm{Hg}(\mathrm{s}, \text { red }))=-90.83 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

⇒ \(\text { Substance } \mathrm{S}^{\ominus}\left(\mathrm{JK}^{-1} \mathrm{~mol}^{-1}\right)\)

⇒ \(\mathrm{Hg}_{\mathrm{g}} 38.01\)

⇒ \(\mathrm{O}_2(\mathrm{~g}) 205.04\)

⇒ \(\mathrm{HgO}(\mathrm{s}, \text { red) } 70.29\)

Which of these reactions islare spontaneous? For those reactions that are not spontaneous,find the temperature at which they will become spontaneous assuming ΔH and εS to be independent of temperature.

Solution:

1. \(\Delta G^{\ominus}=\Delta_f G^{\ominus}\left(\mathrm{Li}_2 \mathrm{CO}_3(\mathrm{~s})\right)+\Delta_{\mathrm{f}} G^{\ominus}\left(\mathrm{H}_2 \mathrm{O}(\mathrm{g})\right)-2 \times \Delta_f G^{\ominus}(\mathrm{LiOH}(\mathrm{s}))-\Delta_f G^{\ominus}\left(\mathrm{CO}_2(\mathrm{~g})\right)\)

= -1132.4 + (-228.6)- 2 x (-443.9)- (-3944)

= -78.8 kj Spontaneous reaction

2. \(\Delta G^{\ominus}=2 \times \Delta_f G^{\ominus}(\operatorname{HBr}(\mathrm{g}))-\Delta_f G^{\ominus}\left(\mathrm{H}_2(\mathrm{~g})\right)-\Delta_f G^{\ominus}\left(\mathrm{Br}_2(\mathrm{l})\right)\)

= 2 x -53.43- 0- 0

= -106.86 kj Spontaneous reaction

3. \(\Delta G^{\ominus}=\Delta_f G^{\ominus}\left(\mathrm{Ca}(\mathrm{OH})_2(\mathrm{~s})\right)-\Delta_{\mathrm{f}} G^{\ominus}(\mathrm{CaO}(\mathrm{s}))-\Delta_f G^{\ominus}\left(\mathrm{H}_2 \mathrm{O}(\mathrm{l})\right)\)

-896.8- (-6042)- (-237.2)

= -55.4 kjSpontaneous reaction

4. \(\Delta G^{\ominus}=2 \times \Delta_f G^{\ominus}(\mathrm{Hg}(\mathrm{l}))+\Delta_f G^{\ominus}\left(\mathrm{O}_2(\mathrm{~g})\right)-2 \times \Delta_f \mathrm{G}^{\ominus}(\mathrm{HgO}(\mathrm{s} ; \text { red }))\)

= 0 + 0 -(-585) = +58.5 kj.

As \(\Delta G^{\ominus}\) is positive, this reaction is not spontaneous. Now let us find \(\Delta H^{\ominus}\) and \(\Delta S^{\ominus}\).

⇒ \(\Delta H^{\ominus}=2 \times \Delta_{\mathrm{f}} H^{\ominus}(\mathrm{Hg}(\mathrm{l}))+\Delta_{\mathrm{f}} H^{\ominus}\left(\mathrm{O}_2(\mathrm{~g})\right)-2 \times \Delta_{\mathrm{f}} H^{\ominus}(\mathrm{HgO}(\mathrm{s}, \mathrm{red}))\)

= 0 + 0- (-90.83) = 90.83 kj.

⇒ \(\Delta S^{\ominus}=2 \times S^{\ominus}(\mathrm{Hg}(\mathrm{l}))+S^{\ominus}\left(\mathrm{O}_2(\mathrm{~g})\right)-2 \times S^{\ominus}(\mathrm{HgO}(\mathrm{s}, \text { red }))\)

= (2 x 38.01) + 205.04- 2 x 70.29 = 140.47 JK-1

For the reaction to be spontaneous T\(\Delta S^{\ominus}\) should exceed \(\Delta H^{\ominus}\). Let us find the temperature at which T\(\Delta S^{\ominus}\) is equal to \(\Delta H^{\ominus}\). Assuming that \(\Delta H^{\ominus}\) and \(\Delta S^{\ominus}\) are independent of temperature.

⇒ T\(\Delta S^{\ominus}\) = \(\Delta H^{\ominus}\)

or T = \(\frac{\Delta H^{\ominus}}{\Delta S^{\ominus}}=\frac{90.83 \times 10^3}{140.47}\) = 646.6 K

Above 646.6 K, the reaction will be spontaneous because then T\(\Delta S^{\ominus}\) > \(\Delta H^{\ominus}\) and \(\Delta G^{\ominus}\) will be negative.

Example 16. Calculate \(\Delta S_295^{\ominus}\) for thefollowing phase changes.

  1. \(\mathrm{C}_6 \mathrm{H}_6(\mathrm{l}) \longrightarrow \mathrm{C}_6 \mathrm{H}_6(\mathrm{~g})\)
  2. \(\mathrm{Cu}(\mathrm{s}) \longrightarrow \mathrm{Cu}(\mathrm{g})\)
  3. \(\mathrm{H}_2 \mathrm{O}(\mathrm{l}) \longrightarrow \mathrm{H}_2 \mathrm{O}(\mathrm{g})\)

The standard entropies of C6H6(l), C6H6(g), Cu(s), Cu(g), H2O(l) and H2O(g) are 172.8, 269.2, 33.15, 1663, 69.91 and 188.71 JK-1 mol-1 respectively.

Solution:

1. \(\Delta S_{\text {vap }}^{\ominus} =S^{\ominus}\left(\mathrm{C}_6 \mathrm{H}_6(\mathrm{~g})\right)-S^{\ominus}\left(\mathrm{C}_6 \mathrm{H}_6(\mathrm{l})\right)\)

=269.2-172.8

∴ \(\Delta S_{\text {vap }}^{\ominus} =96.4 \mathrm{JK}^{-1}\)

2. \(\Delta S_{\text {sub }}^{\ominus} =S^{\ominus}(\mathrm{Cu}(\mathrm{g}))-S^{\ominus}(\mathrm{Cu}(\mathrm{s}))\)

=166.3-33.15

∴ \(\Delta S_{\text {sub }}^{\ominus} =133.15 \mathrm{JK}^{-1}\)

3. \(\Delta S_{\text {vap }}^{\ominus} \left.=S^{\ominus}\left(\mathrm{H}_2 \mathrm{O}(\mathrm{g})\right)-\mathrm{S}^{\ominus} \mathrm{H}_2 \mathrm{O}(\mathrm{l})\right)\)

=188.71-69.91

∴ \(\Delta S_{\text {vap }}^{\ominus} =118.8 \mathrm{JK}^{-1}\) .

Example 17. Calculate the standaril entropy changefor the combustion ofliquid ethanol, C2H5OH to gaseous CO2 and liquid water. The standard entropies of CO2(g), H2O(l), C2H5OH(l) and O2(g) are 213.6 K-1 mol-1, 161JK-1 mol-1 and 205.03 JK-1 mol-1 respectively.
Solution:

⇒ \(\mathrm{C}_2 \mathrm{H}_5 \mathrm{OH}(\mathrm{l})+3 \mathrm{O}_2(\mathrm{~g}) \longrightarrow 2 \mathrm{CO}_2(\mathrm{~g})+3 \mathrm{H}_2 \mathrm{O}(\mathrm{l})\)

⇒  \(\Delta S^{\ominus}=2 \times S^{\ominus}\left(\mathrm{CO}_2(\mathrm{~g})\right)+3 \times S^{\ominus}\left(\mathrm{H}_2 \mathrm{O}(\mathrm{l})\right)-S^{\ominus}\left(\mathrm{C}_2 \mathrm{H}_5 \mathrm{OH}\right)-3 \times\left(S^{\ominus}\left(\mathrm{O}_2(\mathrm{~g})\right)\right.\)

= 2 x 213.4 +3 x 69.91- 161-3 x 205.3 = -139.97 J K-1mol-1.

Example 18. Find the equilibrium constant of the reaction \(\mathrm{CH}_4(\mathrm{~g})+4 \mathrm{Cl}_2(\mathrm{~g}) \rightleftharpoons \mathrm{Cl}_4(\mathrm{~g})+4 \mathrm{HCl}(\mathrm{g})\)

The standard heats offormation of CH4 (g), CCl4 (g) and HCl(g) are -74.81 kJ mol-1,-102.9 kJ mol-1 and -92.31 kJ mol7 respectively. The standard molar entropies of CH4 (g), C2(g), CCl4(g), and HCl(g) are 186.26, 223.07, 309.7 and 186.91 J K-1 mol-1 respectively at 298 K.

Solution:

⇒ \(\Delta_{\mathrm{r}} H^{ominus}=\underset{\text { Products }}{\Sigma \Delta_{\mathrm{f}} H^{ominus}}-\underset{\text { Reactants }}{\Sigma \Delta_f H^{ominus}}\)

= \(\Delta_{\mathrm{f}} H^{\ominus}\left(\mathrm{CCl}_4(\mathrm{~g})\right)+4 \Delta_{\mathrm{f}} H^{\ominus}(\mathrm{HCl}(\mathrm{g}))-\Delta_{\mathrm{f}} H^{\ominus}\left(\mathrm{CH}_4(\mathrm{~g})\right)-4 \Delta_{\mathrm{f}} H^{\ominus}\left(\mathrm{Cl}_2(\mathrm{~g})\right)\)

= -102. 9 + (4 x -92.31)- (-74.81)

Since the standard enthalpies of formation of elements in their reference states are zero by definition

⇒ \(\Delta_{\mathrm{f}} H^{\ominus}\left(\mathrm{Cl}_2(\mathrm{~g})\right) \text { is } 0\)

⇒ \(\Delta_{\mathrm{r}} H^{\ominus} =-397.33 \mathrm{~kJ} \mathrm{~mol}^{-1}\)

∴ \(\Delta_{\mathrm{r} S^{\ominus}} =\underset{\text { Reaction }}{\Sigma S^{\ominus}}-\underset{\text { Products }}{\Sigma S^{\ominus}}\)

= \(S^{\ominus}\left(\mathrm{CCl}_4(\mathrm{~g})\right)+4 S^{\ominus}(\mathrm{HCl}(\mathrm{g}))-S^{\ominus}\left(\mathrm{CH}_4(\mathrm{~g})\right)-4 S^{\ominus}\left(\mathrm{Cl}_2(\mathrm{~g})\right)\)

= 309.7 + (4 x186.91)- 186.26- (4 x 223.07) = -21.2 JK-1 mol-1

Substituting this value of \(\Delta_{\mathrm{r}} S \text { in } \Delta G^{\ominus}=\Delta H^{\ominus}-T \Delta S^{\ominus}\)

⇒ \(\Delta G^{\ominus}\) = -39733 x 103 -298(-21.2) = -391.01 kJ mol-1 = -391.01 kJ mol-1

⇒ \(\Delta G^{\ominus}\) = -RT In K = -2.303 RT log K.

log K = \(-\frac{\Delta G^{\ominus}}{2303 R T}\)

∴ K = \({antilog}\left(\frac{-\Delta G^{\ominus}}{2.303 R T}\right)\)

= \({antilog}\left(\frac{391.01 \times 10^3}{2.303 \times 8.314 \times 298}\right)\)

= antilog (68.53) = 3.37×1068.

Example 19. The standard enthalpy changefor the reaction \(\mathrm{NH}_3(\mathrm{aq})+\mathrm{HCl}(\mathrm{aq}) \longrightarrow \mathrm{NH}_4 \mathrm{Cl}(\mathrm{aq})\) at 298 K is -52.22 kJ. The standard Gibbs energy change is -52.81 kJ. Find the standard molar entropy of NH3(aq) if that of HCl(aq) and NH4Cl(aq) are 56.5 JK-1 mol-1 and 169.9 J K-1 mol-1 respectively.

Solution:

⇒ \(\Delta G^{\ominus}\)=\(\Delta H^{\ominus}-T \Delta S^{\ominus}\)

∴ \(T \Delta S^\ominus=\Delta H^\ominus-\Delta G^\ominus\)

or, \(\Delta S^{\ominus}=\frac{\Delta H^{\ominus}-\Delta G^{\ominus}}{T}\)

= \(\frac{-52220 \mathrm{~J}-(-52810 \mathrm{~J})}{298}=\frac{590}{298}=1.98 \mathrm{~J} \mathrm{~K}^{-1} \text {. }\)

Now \(\Delta S^{\ominus}=S_{\mathrm{NH}_4 \mathrm{Cl}(\mathrm{aq})}^{\ominus}-\left(S_{\mathrm{NH}_3(\mathrm{aq})}^{\ominus}+S_{\mathrm{HCl}(\mathrm{aq})}^{\ominus}\right)\).

Substituting the values, we get

1.098= \(169.9-\left(S_{\mathrm{NH}_3(q)}^{\ominus}+56.5\right)\)

∴ \(S_{\mathrm{NH}_3}^{\ominus}=169.9-56.5-198\)

= \(111.42 \mathrm{JK}^{-1} \mathrm{~mol}^{-1}\).

Example 20. The \(\Delta_{\text {fus }} H^\ominus \text { and } \Delta_{\text {fus }} S^\ominus \text { of } \mathrm{CCl}_4 \text { are } 2.5 \mathrm{~kJ} \mathrm{~mol}^{-1} \text { and } 9.99 \mathrm{~J} \mathrm{~K}^{-1} \mathrm{~mol}^{-1}\) respectively at 298 K. Find the temperature at which solid CCl4 and its liquid are in equilibrium at 1 atm.

Solution:

⇒ \(\Delta G^{\ominus}=\Delta H^{\ominus}-T \Delta S^{\ominus}\).

But at equilibrium, \(\Delta G^{\ominus}\) = 0.

∴ \(\Delta H^{\ominus}\) = T\(\Delta S^{\ominus}\)

or T = \(\frac{\Delta H^{\ominus}}{\Delta S^{\ominus}}\)=\(\frac{2.5 \times 10^3}{9.99}=250.2 \mathrm{~K}\)

Example 21. Consider the reaction \(\mathrm{H}_2(\mathrm{~g})+\mathrm{Cl}_2(\mathrm{~g}) \longrightarrow 2 \mathrm{HCl}(\mathrm{g}).\)

The standard enthalpy change at 298 K is -186.62 kj. The standard molar entropies of H2(g), Cl2(g) and HCl(g) are respectively 130.684, 223.07 and 186.91 J K-1 mol-1 respectively. Calculate the standard Gibbs energy change at 298 Kfor the reaction.

Solution:

⇒ \(\Delta G^{\ominus}\) = \(\Delta H^{\ominus}\) – T \(\Delta S^{\ominus}\)

Given \(\Delta H^{\ominus}\) = -186.62 k] and T = 298 K

∴ \(\Delta S^{\ominus}\) = \(\Delta S^{\ominus}\)(products) – \(\Delta S^{\ominus}\)(reactants)

= \(2 \times S^{\ominus}(\mathrm{HCl}(\mathrm{g}))-S^{\ominus}\left(\mathrm{H}_2(\mathrm{~g})\right)-S^{\ominus}\left(\mathrm{Cl}_2(\mathrm{~g})\right)\)

= 2 x 186.91- 130.684- 223.07 = 20.07 JK-1.

Substituting this value of \(\Delta S^{\ominus}\), we get

⇒ \(\Delta G^{\ominus}\) = -186.62 x 103 – 298 x 20.07

= -186,620-5980.86 = -192,600.86 J.

The standard free energy change for the reaction is -192.6 kj.

Example 22. Predict whether the entropy of the following reactions wll increase or decrease when the reaction occurs.

  1. \(\mathrm{CaSO}_4 \cdot 2 \mathrm{H}_2 \mathrm{O}(\mathrm{s}) \longrightarrow \mathrm{CaSO}_4(\mathrm{~s})+2 \mathrm{H}_2 \mathrm{O}(\mathrm{g})\)
  2. \(\mathrm{CH}_4(\mathrm{~g})+\mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{C}(\mathrm{s})+2 \mathrm{H}_2 \mathrm{O}(\mathrm{g})\)
  3. \(\mathrm{Cu}(\mathrm{s})+\mathrm{S}(\mathrm{g}) \longrightarrow \mathrm{CuS}(\mathrm{s})\)
  4. \(\mathrm{N}_2(\mathrm{~g})+2 \mathrm{O}_2(\mathrm{~g}) \longrightarrow 2 \mathrm{NO}(\mathrm{g})\) .

Solution:

Verify your predictions by calculating \(\Delta S^{\ominus}\) using \(S^{\ominus}\) values.

1. In the reactant side we have only a solid whereas in the product side, we also have a gas apart from a solid. Since gases have large entropy, the entropy of system increases.

∴ \(\Delta S^{\ominus}=S^{\ominus}\left(\mathrm{CaSO}_4(\mathrm{~s})+2 \times S^{\ominus}\left(\mathrm{H}_2 \mathrm{O}(\mathrm{g})\right)-S^{\ominus}\left(\mathrm{CaSO}_4 \cdot 2 \mathrm{H}_2 \mathrm{O}\right)\right.\)

= 106.5 + 2 x 188.71-194.0 = 289.92 JK-1

2. The number of gas molecules (two) is the same on both the reactant and the product side.

Depending on the actual entropies of the substances, \(\Delta S^{\ominus}\) may increase or decrease but only slightly.

∴ \(\Delta S^{\ominus}=S^{\ominus}(\mathrm{C}(\mathrm{s}))+2 \times S^{\ominus}\left(\mathrm{H}_2 \mathrm{O}(\mathrm{g})\right)-S^{\ominus}\left(\mathrm{CH}_4(\mathrm{~g})\right)-S^{\ominus}\left(\mathrm{O}_2(\mathrm{~g})\right)\)

= 5.74 + 2 x188.71- 186.15- 205.03 = -8.02JK-1

3. One of the reactants is in the gaseous state but the only product is a solid. Hence, the entropy will decrease.

∴ \(\Delta S^{\ominus}=S^{\ominus}(\mathrm{CuS}(\mathrm{s}))-S^{\ominus}(\mathrm{Cu}(\mathrm{s}))-S^{\ominus}(\mathrm{S}(\mathrm{g}))\)

= 66.5- 33.15-167.75 =1344 J K-1.

4. The total number of gas molecules decreases from three to two as the reaction proceeds from left to right. This results in a decrease in entropy.

∴ \(\Delta S^{\ominus}=2 \times S^{\ominus}(\mathrm{NO}(\mathrm{g}))-S^{\ominus}\left(\mathrm{N}_2(\mathrm{~g})\right)-2 \times S^{\ominus}\left(\mathrm{O}_2(\mathrm{~g})\right)\)

= 2 x 210.65-1915- 2 x 205.03 = -180.26 JK-1

Thermodynamics Multiple Choice Questions

Question 1. A well-stoppered thermos flask containing some ice cubes is an example of a

  1. Closed system
  2. Open system
  3. Isolated system
  4. Nonthermodynamic system

Answer: 3. Isolated system

Question 2. Which of the following is correct in the context of an adiabatic process?

  1. pΔV = 0
  2. q = +W
  3. ΔU=q
  4. q = 0

Answer: 4. q = 0

Question 3. The enthalpy change for the reaction \(\mathrm{C}(\mathrm{s})+\mathrm{O}_2(\mathrm{~g}) \rightarrow \mathrm{CO}_2(\mathrm{~g})\) is

  1. Positive
  2. Negative
  3. Zero
  4. None of these

Answer: 2. Negative

Question 4. The enthalpy of formation of ammonia is – 46.0 kJ mol-1 What is the enthalpy change for the following reaction?

\(2 \mathrm{NH}_3(\mathrm{~g}) \longrightarrow \mathrm{N}_2(\mathrm{~g})+3 \mathrm{H}_2(\mathrm{~g})\)

  1. + 46.0 kj mol-1
  2. – 23.0 kj mol-1
  3. – 92.0 kj mol-1
  4. + 92.0 kj mol-1

Answer: 4. + 92.0 kj mol-1

Question 5. The heat released when 0.6 mol of HNO3 solution is mixed with 0.2 mol of NaOH is

  1. 57.0 kj
  2. 11.4 kj
  3. 28.5kj
  4. 34.9 kj

Answer: 2. 11.4 kj

Question 6. When NaCl dissolves in water, the entropy

  1. Increases
  2. Decreases
  3. Remains the same
  4. None of these

Answer: 1. Increases

Question 7. The enthalpy change for the reaction \(\mathrm{H}_2(\mathrm{~g})+\mathrm{1/2}\mathrm{O}_2(\mathrm{~g}) \longrightarrow \mathrm{H}_2 \mathrm{O}(\mathrm{g})\) is called

  1. Enthalpy of formation of water
  2. Enthalpy of combustion of hydrogen
  3. Enthalpy of vaporisation of water
  4. None of these

Answer: 1. Enthalpy of formation of water and 2. Enthalpy of combustion of hydrogen

Question 8. The energy required to dissociate 4 g of gaseous hydrogen into free gaseous atoms is 208 kcal at 25°C. The bond energy of the H—H bond is

  1. 104 kcal mol-1
  2. 10.4 kcal mol-1
  3. 208 kcal mol-1
  4. 416 kcal mol-1

Answer: 1. 104 kcal mol-1

Question 9. In which of the following does entropy decrease?

  1. Crystallisation of sugar from a solution
  2. Rusting of iron
  3. Melting of ice
  4. Vaporisation of camphor

Answer: 1. Crystallisation of sugar from a solution

Question 10. For the reaction X→y Y, ΔH and ΔS are negative at a certain temperature. This reaction will be spontaneous at

  1. Low temperature
  2. High temperature
  3. The same temperature
  4. Very high temperature

Answer: 1. Low temperature

Question 11. A spontaneous reaction is impossible if

  1. Both ΔH and ΔS are positive
  2. ΔH and ΔS are negative
  3. ΔH is positive and ΔS is negative
  4. ΔH is negative and ΔS is positive

Answer: 3. ΔH is positive and ΔS is negative

Question 12. When a solid melts, there is

  1. An increase in entropy
  2. An increase in enthalpy
  3. A decrease in internal energy
  4. Decrease in enthalpy

Answer: 1. An increase in entropy and 2. An increase in enthalpy

Question 13. Which of the following conditions are favourable for a spontaneous process?

  1. ΔH=-ve TΔS = + ve
  2. ΔH = -ve TΔS = -ve TΔS<ΔH
  3. ΔH=+ve TΔS = + ve TΔS<ΔH
  4. ΔH = +ve TΔS = + ve TΔS>ΔH

Answer: 1. ΔH=-ve TΔS = + ve,

2. ΔH = -ve TΔS = -ve TΔS<ΔH and

4. ΔH = +ve 7ΔS = + ve TΔS>ΔH

Question 14. The enthalpies of elements in their standard states are taken as zero. Thus, the enthalpy of formation of a compound

  1. Will always be positive
  2. Will always be negative
  3. Will always be zero
  4. May be negative or positive

Answer: 4. May be negative or positive

Question 15. The enthalpy of combustion of a substance

  1. Is always positive
  2. Is always negative
  3. Is numerically equal to the enthalpy of formation
  4. Cannot be predicted

Answer: 2. Is always negative

Question 16. When ammonium chloride is dissolved in water, the solution becomes cold. The change is

  1. Exothermic
  2. Endothermic
  3. Super cooling
  4. None of these

Answer: 2. Endothermic

WBBSE Solutions For Class 6 Maths Chapter 25 Fun With Numbers

Class 6 WBBSE Math Solutions Chapter 25 Fun With Numbers Exercise 25

Question 1. Let’s observe the interesting numbers.
Solution:

Let’s observe the interesting numbers

11 x 11 = 121

11 x 11 x 11 = 1331

11 x 11 x 11 x11 = 14641

11 x 11 x 11 x 11 x 11 = 161051

Other interesting numbers

10 x 1 = 10; 1+0=1

11 x 1 = 11; 1+1=2

12 x 1 = 12; 1+2=3

13 x 1 = 13; 1+3=4

14 x 1 = 14; 1+4=5

15 x 1 = 15; 1+5=6

16 x 1 = 16; 1+6=7

17 x 1 = 17; 1+7=8

18 x 1 = 18; 1+8=9

19 x 1 = 19; 1+9=10

Read and Learn More WBBSE Solutions For Class 6 Maths

An interesting arrangement of the square and square root of natural number.

1 =1 =1²

1 + 3 = 4 = 2²

1 + 3 + 5 = 9 = 3²

1 + 3 + 5 + 7 = 16 = 4²

1 + 3 + 5 + 7 + 9 = 25 = 5²

1 + 3 + 5 + 7 + 9 + 11 = 36 = 6²

1 + 3 + 5 + 7 + 9 + 11 + 13 = 49 = 7²

Class 6 WBBSE Math Solutions

Question 2. By arranging the buttons, let’s understand the relation between square and square root of natural numbers.
Solution:

By arranging the buttons, let’s understand the relation between square and square root of natural numbers.

WBBSE Solutions For Class 6 Maths Chapter 25 Fun With Numbers Square And Square Root Of Natural Numbers

Class 6 Math Solutions WBBSE Chapter 25 Fun With Numbers Exercise 25.1

Question 1. With four 4 ‘4s’, let us make numbers from 1 to 18.
Solution:

1. \(\frac{4+4}{4+4}=\frac{8}{8}=1\)

2. \(\frac{4}{4}+\frac{4}{4}=1+1=2\)

3. \(\frac{4+4+4}{4}=\frac{12}{4}=3\)

4. \(\frac{4-4}{4}+4=0+4=4\)

5. \(\frac{4+4 \times 4}{4}=-\frac{4+16}{4}=\frac{20}{4}=5\)

6. \(4+\frac{4+4}{4}=4+2=6\)

7. \(4+4-\frac{4}{4}=8-1=7\)

8. \(4+\frac{4 \times 4}{4:}=4+4=8\)

9. \(4+4+\frac{4}{4}=8+1=9\)

10. \(\frac{44-4}{4}=\frac{40}{4}=10\)

11. \(\frac{4}{4}+\frac{4}{4}=\frac{4 \times 10}{4}+1=11\)

12. \(\frac{44+4}{4}=\frac{48}{4}=12\)

13. \(\frac{44}{4}+\sqrt{4}=11+2=13\)

14. \(4(4-4)-4=14 \cdot 4-4=14\)

15. \(4 \times 4-\frac{4}{4}=16-1=15\)

16. \(4+4+4+4=16\)

17. \(4 \times 4+\frac{4}{4}=16+1=17\)

18. \(44 x \cdot 4+\cdot 4=17 \cdot 6+\cdot 4=18\)

Question 2. With five ‘9s’, let me make 1000.
Solution:

⇒ \(999+\frac{9}{9}=999+1000\)

Question 3. Let’s solve puzzles with Roman letters:

1. If 9 is taken from 6,10 is taken from 9,50 is taken from 40, we get 6 – let’s try now it can be made possible.
Solution:

If 9 is taken from 6 i.e., VI-IX = 0,

10 is taken from 9 i.e., IX – X = 1

50 is taken 40 i.e., XL – L = 2

∴ We get, with three match sticks I made VI.

2. By just shifting one match stick, let us get our answer correct.
Solution: As there are three sticks on both sides.

WBBSE Class 6 Maths Solutions

Question 4. There are 5 apples in a basket. Let’s divide them among 5 girls in such a way that each gets 1 apple but one apple remains in the basket.
Solution: If 4 girls get I apple each while the fifth girl gets the basket in the remaining apple still in it.

Question 5. Azim has 3 match sticks. He asked me to write 4 with those 3 match sticks. Let me try to do.
Solution: Azim has 3 match sticks I write 4 with 3 match sticks IV.

Question 6. I have 3 match sticks. I took 2 more sticks to write 8.
Solution: I have 3 match sticks. I took 2 more sticks.

Total number of sticks = 3 + 2 + 5

I write – VIII.

Question 7. I have 7 match sticks and 6 buttons. I arranged the 7 match sticks on a table in a form given below.
Solution:

If I start from 1, the next one will be 3.

I start with 2, the next one will be 4.

I start at 5, the next one will be 7.

I start from 6, and the next one will be 1.

The button I have placed already.

So, next can not start from button 6.

Question 8. Few other fun with match sticks:

1. With 6 match sticks I made 1/7 as WBBSE Solutions For Class 6 Maths Chapter 25 Fun With Numbers
Solution:

With 5 match sticks I made \(\frac{1}{3} \text { as } \frac{1}{\text { III }}\)

WBBSE Class 6 Maths Solutions

2. Let’s make 2 fractions whose value is 1/3.
Solution:

With 6 match sticks I made \(\frac{2}{6} \text { as } \frac{11}{V 1}\)

Question 9. I have 12 match sticks. I shall make a puzzle with these 12 match sticks and try to solve them. With these 12 sticks, 4 squares (and a big square) have been found.

1. First removing 2 sticks, 2 different-sized squares are made.
Solution:

WBBSE Solutions For Class 6 Maths Chapter 25 Fun With Numbers One Small Squares

One small and one big squares.

WBBSE Solutions For Class 6 Maths Chapter 25 Fun With Numbers One Big Squares

2. By shifting 3 sticks, let’s try to get 3 equal-sized squares.
Solution:

WBBSE Solutions For Class 6 Maths Chapter 25 Fun With Numbers Three Equal Sized Squares

Three equal-sized squares

3. Again, by changing side of 4 sticks, 3 equal-sized squares can be formed
Solution:

WBBSE Solutions For Class 6 Maths Chapter 25 Fun With Numbers Three Equal Sized Squares (2)

Three equal-sized squares.

4. Just by moving 2 sticks, let’s try to get 7 different-sized squares. (In this case 2 sticks may be placed diagonally.)
Solution:

WBBSE Solutions For Class 6 Maths Chapter 25 Fun With Numbers Seven Different Sized Squares

Seven different-sized squares.

WBBSE Math Solutions Class 6

5. I moving 4 sticks, let’s try to make 10 small and big squares (but in this case 2 sticks can be put diagonally).
Solution:

WBBSE Solutions For Class 6 Maths Chapter 25 Fun With Numbers 10 Small And Big Squares

10 small and big squares.

Question 10. We know 2 + 2 = 2 x 2 i.e., two 2’s added and two 2’s multiplied gives the same value.

1. For 3 natural numbers _____, _____, and _____, it is found that their sum is equal to their product. i.e., ____ + ______ + _______ = ______ x _______ x ________
Solution: 1 + 2 + 3 = 1 x 2 x 3

2. Let us find 4 natural numbers, whose sum and product are the same. i.e., ____ + ____ + ______+ ____ + ______ + _____ + _______ + _____
Solution: 1+1+2+4=1x1x2x4

3. Is it possible to have 5 natural numbers where this relation holds good ? (find myself) Yes, I can make the sum and product of 5 natural numbers are equal, one example is: 1+1+1+2+5=1x1x1x2x5

Let us try to find other numbers _____ + _____ + _____ + _____ + ______ = ______ x _______ x _______ x ________ x ________

Solution: 1 + 1 + 1 + 3 + 3 = 1 x 1 x 1 x 3 x 3

Question 11.

1 x 1 = 1

11 x 11 = 121

111 x 111 = 12321

1111 x 1111 = 1234321

_______ x _____ = ________

_________ x ______ = _______

Solution:

1 x 1 = 1

11 x 11 = 121

111 x 111 = 12321

1111 X 1111 = 1234321

11111 X 11111 = 123454321

111111 X 111111 = 12345654321

 

WBBSE Solutions For Class 6 Maths Chapter 27 Equivalence Of Fraction Decimal Fraction Percentage And Ratio

Class 6 Math Solutions WBBSE Chapter 27 Equivalence Of Fraction Decimal Fraction Percentage And Ratio Exercise 27

Question 1. Today we have cut off different geometric-shaped pieces of paper. Preetam took circular and rectangular pieces. He then divided these circular and rectangular pieces into equal parts and colored different parts.
Solution:

Now, let’s write without looking at the illustration:

Read and Learn More WBBSE Solutions For Class 6 Maths

WBBSE Solutions For Class 6 Maths Chapter 27 Equivalence Of Fraction Decimal Fraction Percentage And Ratio Values Illustartion

WBBSE Solutions For Class 6 Maths Chapter 26 Open Shapes Of Regular Solids

Class 6 Math Solutions WBBSE Chapter 26 Open Shapes Of Regular Solids Exercise 26

Question 1. I opened a cube-shaped paper box.

WBBSE Solutions For Class 6 Maths Chapter 26 Open Shapes Of Regular Solids Cube Shaped Paper Box

Which of the following shapes shall I get?

WBBSE Solutions For Class 6 Maths Chapter 26 Open Shapes Of Regular Solids 3 Is Cube Shaped Paper Box

Solution: 3

Question 2. I opened a cuboid-shaped paper box.

Read and Learn More WBBSE Solutions For Class 6 Maths

WBBSE Solutions For Class 6 Maths Chapter 26 Open Shapes Of Regular Solids A Cuboid Shaped Paper Box

Which of the following shapes shall I get?

WBBSE Solutions For Class 6 Maths Chapter 26 Open Shapes Of Regular Solids 1 Is A Cuboid Shaped Paper Box

Solution: 1

Question 3. A paper-made tetrahedron is opened.

WBBSE Solutions For Class 6 Maths Chapter 26 Open Shapes Of Regular Solids Paper Made Terahedron Is Opened

Let’s find which shape we will get.

Class 6 WBBSE Math Solutions

WBBSE Solutions For Class 6 Maths Chapter 26 Open Shapes Of Regular Solids 2 Is A Paper Made Terahedron Is Opened

Solution: 2

Question 4. Now I opened this soild.

WBBSE Solutions For Class 6 Maths Chapter 26 Open Shapes Of Regular Solids Opened This Solid

Let’s find the shape.

WBBSE Solutions For Class 6 Maths Chapter 26 Open Shapes Of Regular Solids A Is Opened This Solid

Solution: 1

Question 5. I opened this solid.

WBBSE Solutions For Class 6 Maths Chapter 26 Open Shapes Of Regular Solids Open Solid

Which shape will it be?

WBBSE Solutions For Class 6 Maths Chapter 26 Open Shapes Of Regular Solids 3 Is A Open Solid

Solution: 3

WBBSE Math Solutions Class 6

Question 6. Let’s open the solid

WBBSE Solutions For Class 6 Maths Chapter 26 Open Shapes Of Regular Solids Open Solid 1

Let us find the shape.

WBBSE Solutions For Class 6 Maths Chapter 26 Open Shapes Of Regular Solids 3 Is A Open Solid 1

Solution: 3

Question 7. This solid is opened.

WBBSE Solutions For Class 6 Maths Chapter 26 Open Shapes Of Regular Solids Open Solid 2

The shape will be

WBBSE Solutions For Class 6 Maths Chapter 26 Open Shapes Of Regular Solids 3 Is Open Solid 2

Solution: 4

 

WBBSE Solutions For Class 6 Maths Chapter 24 Solids From Different Sides Perspective

Class 6 Math Solutions WBBSE Chapter 24 Solids From Different Sides Perspective Exercise 24

Question 1. I have many plastic cubes. By putting them together different solid shapes can be formed. Let me find out how these shapes will look when viewed from front, side, and 1rom top. Let us write in the blank space how these solids will look [from side/from top].

WBBSE Solutions For Class 6 Maths Chapter 24 Solids From Different Sides Perspective Plastic Cubes Putting Them Together Different Solid Shapes

Solution:

Read and Learn More WBBSE Solutions For Class 6 Maths

WBBSE Solutions For Class 6 Maths Chapter 24 Solids From Different Sides Perspective Shapes Will Look When Viewed From Front Side And From Top

WBBSE Solutions For Class 6 Maths Chapter 23 Symmetry

Class 6 Math Solutions WBBSE Chapter 23 Symmetry Exercise 23

In the pictures given below, let us find the pictures which are not symmetrical and let’s colour them, and the pictures which are symmetrical, let’s mark the line/lines of symmetry.
Solution:

WBBSE Solutions For Class 6 Maths Chapter 23 Symmetry Symmetric And Non Symmetric Pictures

Question 1. A graph paper with square cells, taking the straight line AB as a line of symmetry, let us complete the symmetrical pictures.
Solution:

WBBSE Solutions For Class 6 Maths Chapter 23 Symmetry A Graph Paper With Square Cells Taking Straigth Line AB

Question 2. Let me draw 5 symmetrical pictures I have seen.
Solution:

Read and Learn More WBBSE Solutions For Class 6 Maths

WBBSE Solutions For Class 6 Maths Chapter 23 Symmetry 5 Symmetrical Pictures

Question 3. In the figure given, taking AB and CD as the lines of symmetry, let us draw a symmetrical picture.
Solution:

WBBSE Solutions For Class 6 Maths Chapter 23 Symmetry Taking AB And CD As Line Of Symmetry

Question 4. In the pictures on the graph paper given below, let’s draw the line/lines of symmetry and note down the number of lines of symmetry that are possible for each figure.

Class 6 Math Solutions WBBSE
Solution:

WBBSE Solutions For Class 6 Maths Chapter 23 Symmetry Lines Of Symmetry

Question 5. Using graph paper let’s draw.

1. A triangle having 1 line of symmetry.
Solution:

WBBSE Solutions For Class 6 Maths Chapter 23 Symmetry A Triangle Having 1 Line Of Symmetry

2. A triangle having no line of symmetry.
Solution:

WBBSE Solutions For Class 6 Maths Chapter 23 Symmetry A Triangle Having No Line Of Symmetry

3. A triangle having 3 lines of symmetry.
Solution:

Class 6 Math Solutions WBBSE

WBBSE Solutions For Class 6 Maths Chapter 23 Symmetry A Triangle Having 3 Lines Of Symmetry

4. A quadrilateral having 4 lines of symmetry.
Solution:

WBBSE Solutions For Class 6 Maths Chapter 23 Symmetry A Quadrilateral Having 4 Lines Of Symmetry

5. A quadrilateral has 2 lines of symmetry.
Solution:

WBBSE Solutions For Class 6 Maths Chapter 23 Symmetry A Quadrilaterla having 2 Lines Of Symmetry

6. A quadrilateral having just 1 line of symmetry.
Solution:

WBBSE Solutions For Class 6 Maths Chapter 23 Symmetry A Quadrilateral Having Just 1 Lines Of Symmetry

7. A quadrilateral which does not have a line of symmetry.
Solution:

WBBSE Solutions For Class 6 Maths Chapter 23 Symmetry A Qualdrilateral Which Does Not Have A Line Of Symmetry

Question 6. Let’s draw in the exercise book and mention the number of line of symmetry.
Solution:

WBBSE Solutions For Class 6 Maths Chapter 23 Symmetry Number Of Lines Of Symmetry

Question 7. Let’s draw the reflection on the mirror.
Solution:

Class 6 Math Solutions WBBSE

WBBSE Solutions For Class 6 Maths Chapter 23 Symmetry Reflection On The Mirror

Question 8. Let’s draw the lines of symmetry for the following pictures.
Solution:

WBBSE Solutions For Class 6 Maths Chapter 23 Symmetry Draw Lines Of Symmetry

WBBSE Solutions For Class 6 Maths Chapter 22 Drawing Of Different Geometrical Figures

Class 6 WBBSE Math Solutions Chapter 22 Drawing Of Different Geometrical Figures Exercise 22

Question 1. In the following figures, let’s identify which two straight lines have intersected at right angles.
Solution:

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures Two Straigth Lines Have Intersected At Rigth Angles

In the figures (2) and (3), the two straight lines intersected at right angles.

Question 2. Let’s fill in the blanks:

1. At 3’o clock the arms of the clock will be at _____
Solution: A 3’o clock the arms of the clock will be at [right angles].

2. The straight lamp post on the road is at _______ to the ground.
Solution: The straight lamp post on the road is at [right angles] to the ground.

Question 3. Let me write two such things which are at right angles to each other.
Solution: Walls and the floor of a room are at right angles to each other.

Read and Learn More WBBSE Solutions For Class 6 Maths

Question 4. Let us draw a line segment AB of length 4 cm; a point O is taken on the line segment AB, such that AO = 1 cm and OB = 3 cm. Now, at O, let’s draw a perpendicular AB at O.
Solution:

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures Two Line Segments Perpendicular AB At O

Question 5. A line segment PQ of length 4 cm is drawn. A point O is taken on the segment PQ. With the help of scale and pencil, compass, let’s draw perpendicular OM on PQ at O.

Class 6 WBBSE Math Solutions
Solution:

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures A Line Segments Perpendicular OM On PQ At O

Draw a line segment PQ of 4 cm. Take a point O on PQ. Now, a perpendicular OR is drawn on PQ at O.

Question 6. A line segment XY of length 5 cm is drawn. A point P is taken outside the line segment XY. With a set square, let’s draw a perpendicular PL, on line segment XY from the point P.
Solution:

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures A Line Segments XY From The Point P

A line segment XY of length 5 cm is drawn. A point is taken outside the line segment XY.

Now, with the help of a scale and a set square, I draw a perpendicular PL on the line segment XY from the point P.

Class 6 WBBSE Math Solutions

Question 7. A line segment AB of length 8 cm is taken. A point P is taken on the line segment AB, so that AP = 3 cm, and BP = 5 cm, let’s draw a perpendicular PL on the line segment AB, at point P with the help of scale and pencil and a compass.
Solution:

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures Perpendicular PL On The Line Segment AB At Point P

Draw a straight line AB of 8 cm. A point P is taken on AB, such that AP – 3 cm and PB = 5 cm.

Now, a perpendicular PL is drawn on AB at P with scale pencil and a compass.

Question 8. A line segment AB of length 6 cm is drawn. Let’s take a point K outside this line segment. With the help of scale and compass, let’s draw a perpendicular KL on the line segment AB.
Solution:

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures Draw A Perpendicular KL On The Line Segment AB

A straight line AB of length 6 cm is taken. A point K is taken outside AB. Now, with centre K, an arc is drawn which cuts AB at P and Q.

Now, with centre P and Q, two arcs with radii more than of 1/2 of PQ are drawn, which cut each other at R. Join KR which cuts AB at L. KL is the required perpendicular on AB.

Class 6 WBBSE Math Solutions

Question 9. Let’s draw a triangle ABC with the help of scale and pencil. From the three vertices A, B and C of the triangle, perpendiculars AP, BQ and CR are drawn on the sides BC, AC and AB respectively. Let’s see from the figure, if the line segments AP, BQ and CR are concurrent, i.e., they intersect at a point.
Solution:

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures Three Perpendicular AP BQ And CR Meet At O

A triangle ABC is drawn. From three vertices A, B and C of the triangle ABC, three perpendiculars AP, BQ and CR are drawn. The three perpendiculars AP, BQ and CR meet at O, i.e., perpendiculars are concurrent.

Class 6 Maths Solutions WBBSE

Question 10. With the help of a set square, let us draw a right-angled triangle ABC such that ABC = ∠90°. From point B, let’s draw a perpendicular on the hypotenuse AC. Let’s also find if the perpendiculars from A, B and C on the opposite sides meet at point. If so, name the point.
Solution:

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures Perpendicular From A B And C On The Opposite Side Meet At Point

ABC is a right-angled triangle, ABC = ∠90°. A perpendicular BD is drawn on AC. AB is perpendicular on BC from point A and BC is particular on AB from point C.

These three perpendiculars meet at B.

Question 11. At a fixed point on a straight line ___ perpendicular can be drawn.
Solution: At a fixed point on a straight line one perpendicular can be drawn.

Question 12. From a fixed point outside a straight line _____ (one/more than one) perpendicular can be drawn to that straight line.
Solution: From a fixed point outside a straight line one perpendicular can be drawn to the straight line.

Class 6 Maths Solutions WBBSE

Question 13. Let’s draw a line segment AB, let’s take any two points P and Q on the line segment. Let’s draw two perpendiculars PM and QN at P and Q on the line segment AB; let’s find if line segments PM and QN are parallel or intersecting.
Solution:

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures Line Segments PM And QN Are Parallel Or Intersecting

Let P and Q are the two points on the line segment AB. Two perpendiculars r PM and QN are drawn on AB at P and Q. Here PM and QN are two parallel lines.

Question 14. Let’s draw a line segment of length 4 cm; taking this length as radius a circle is drawn. Let the centre of the circle be named O. A chord AB which is not a diameter is drawn. From O a perpendicular OM is drawn on the chord AB. Let’s measure the line segments AM and BM with a scale and write the conclusion derived.
Solution:

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures Line Segments AM And BM With A Scale

O is the centre of the circle of radius 4 cm (OC).

AB is a chord. A perpendicular OM is drawn from the centre O. AB is bisected at M.

AB = 6.6 cm; AM = BM = 3.3 cm.

Class 6 Maths Solutions WBBSE Chapter 22 Drawing Of Different Geometrical Figures Exercise 22.1

I drew a line segment XY with scale and pencil on my exercise book. I Now we shall draw the perpendicular bisector of line segment XY with the help of scale, pencil and pencil compass.
Solution:

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures Perpendicular Bisector Of Line Segment XY

A line segment XY is drawn with a scale and pencil. Now, with centres X and Y, two arcs with the same radius are drawn on both sides of XY.

The arcs intersect each other at C and D. Join C and, D which cuts XY at P.

∴ XP = PY = 4 cm.

Question 1. If taken a radius, less than half the length of the line segment XY and try to draw an arc with centres X and Y, on both sides, what shall I find?
Solution: Now, with centres X and Y, we draw two arcs with a radius less than half the length of line segment XY on both sides.

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures Two Arcs With Radius Less Than Half The Length Of Linex XY On Both Sides

Question 2. Again, if I draw arcs with centers X and Y and radius equal to the length 1 more than half the length of the line segment, what shall I observe?
Solution:

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures Two Arcs Cut The Previous Arcs At QR And MN

Again, with centres X and Y, we draw two more arcs with a radius half of the length of line segment XY.

The two arcs cut the previous arcs at QR and MN.

WBBSE Math Solutions Class 6 Chapter 22 Drawing Of Different Geometrical Figures Exercise 22.2

Question 1. Let’s draw a line segment AB of length 5 cm with scale and pencil. By folding the paper, let us find the perpendicular bisector of the line segment AB. Then, we actually measure the two bisected parts with scale.
Solution:

1. Let’s take a thick paper. It is folded at any end and then unfolded. The line formed by the fold is marked with a pencil.

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures The Line Formed By The Fold Is Marked With A Pencil

2. On this, two points A and B are taken. Now the paper is again folded in such a way that points A and B coincide. Again, on this fold mark a line segment. CD is drawn and we get the mid point of the line segment AB. That D is the midpoint of the line segment AB and the line segment CD is perpendicular at D on line segment AB.

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures Line Segments CD Is Perpendicular To The Line Segments AB At Mid Point

Hence, the line segment CD is perpendicular to the line segment AB at its midpoint. Line segment CD is called the Perpendicular Bisector of the line segment AB.

The perpendicular bisector of a line segment divides it into two equal parts.

WBBSE Math Solutions Class 6

Question 2. Let’s draw a line segment of length 8 cm with scale and pencil. Then let’s draw a perpendicular bisector of this line segment and each part measured.
Solution:

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures Perpendicular Bisector Of A Line Segment Divides Into 2 Equal Parts

Let AB is a line segment of 8 cm. Now with centres A and B two arcs with radii more than 1/2 of AB are drawn on both sides of AB. They intersect each other at C and D. Join C, D which cuts AB at P.

AP = PB = 4 cm.

Question 3. Let’s draw a line segment PQ of length 6 cm with scale and pencil, now, taking line segment PQ as diameter, let’s draw a circle.
Solution:

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures PQ As Diameter To A Circle

Let PQ is a line segment of length 6 cm. It is bisected at O. Now, with centre O and radius PO a circle is drawn.

WBBSE Math Solutions Class 6

Question 4. With the help of scale and pencil, a line segment AB of length 8 cm is drawn. Let’s divide the line segment AB into four equal parts and measure each part.
Solution:

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures Divides The Line Segements AB Into Four Equal parts

  1. First we bisect the line segment AB into equal parts such that AR = RB.
  2. Again, we bisect the line segment AR and RB into two equal parts such that AE = ER and RF = FB. Hence, the segment AB is divided four equal parts.

∴ AE = EF = RF = FB = 2 cm.

WBBSE Math Solutions Class 6

Question 5. Let’s draw two circles whose diameters are 5 cm and 7 cm respectively.
Solution:

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures Two Circles Diameters Are 5 cm And 7 cm

AB and PQ are two line segments of lengths 5 cm and 7 cm respectively. Line segment AB is bisected at C and line segment PQ is bisected at R.

Now, with centres C and R two circles are drawn with radii of 2.5 cm and 3.5 cm respectively.

Question 6. Masum drew a triangle ABC. Then with the help of scale and pencil, compass, she bisected the sides BC, AC and AB perpendicularly. She verified that the three perpendicular bisectors are concurrent. When found concurrent, she named it O. Now with centre O and radius equal to the line segment AO, she drew a circle.
Solution:

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures Three Perpendicular AP BQ And CR Meet At O

ABC is a triangle, the sides BC, CD and DA are bisected at D, E, and F. These 3 bisectors meet at O. Now, with centre O and radius OA, a circle is drawn which passes through B and C. Point O is called circumcentre of the triangle ABC.

WBBSE Class 6 Maths Solutions Chapter 22 Drawing Of Different Geometrical Figures Exercise 22.3

Question 1. Let us draw the following angles with the help of protractor: 30°, 42°, 105°, 67°, 88°, 120°, 205°, 282°.
Solution:

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures Angles With Protractor

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures Angles With Protractor-1

Question 2. When a clock shows the following time, the angles formed by their arms are to be drawn with a protractor on the exercise book.

1. 3 p.m.
Solution: At 3 p.m.: The angle formed by the arms is 90°.

2. 5 a.m.
Solution: At 5 p.m.: The angle formed by the arms is 150°.

3. 10a.m
Solution: At 10 a.m.: The angle formed by the arms is 60°.

4. 4 p.m.
Solution: At 4 p.m.: The angle formed by the arms is 120°.

WBBSE Class 6 Maths Solutions

Question 3. Let’s draw the following angles with the help of set square and bisect them with a scale, pencil and compass.

1. 30°
Solution:

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures Angle ABC Is 30 Degrees

∠ABC = 30°

PB is the bisector ∠ABC.

2. 45°
Solution:

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures Angle PQR Is 45 Degrees

∠PQR = 45°

QC is the bisector ∠PQR.

3. 60°
Solution:

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures Angle MBN Is 60 Degrees

∠MBN = 60

BR is the bisector of ∠MBN.

4. 90°
Solution:

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures Angle XYZ Is 90 Degrees

∠XYZ = 90°

YP is the bisector of ∠XYZ.

5. 105°
Solution:

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures Angle ABC Is 105 Degrees

∠ABC = V ABR + VRBC = 60° + 45° = 105°

∠BK is the bisector of ∠ABC.

WBBSE Class 6 Maths Solutions

Question 4. Without using a protractor let us draw an angle of 45° with a scale, pencil, compass.
Solution:

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures Angle 45 Degrees Using Protractor With Scale Pencil Compass

1st BAC = 90° is drawn

Now, ∠BAC is bisected, ∠BAD = 45°, and ∠CAD = 45°.

Question 5. Let’s draw an angle of 120° with the help of a protractor, then divide it into 4 equal parts with the help of scale, pencil and compass.
Solution:

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures Angle Of 120 Degrees With Help Of Protractor

With the protractor 1st ∠PQR = 120° is drawn. Now, 1st bisect the ∠PQR into two equal parts.

QK is the bisector of ∠PQR. The two equal angles are ∠PQK and ∠KQR, each equal to 60°.

Again, bisect two angles ∠PQK and ∠KQR.

∴ 4 equal parts of ∠PQR are ∠PQM = ∠MQK = ∠KQN = ∠NQR = 30°.

WBBSE Class 6 Maths Solutions

Question 6. Let’s draw a triangle ABC with the help of a scale and pencil. Then the three angles of the triangle are bisected. Let’s find if the bisectors of the angles are concurrent.
Solution:

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures A Triangle ABC With Help Of A Scale And Pencil

Let ABC is a triangle. Its 3 angles are ∠BAC, ∠ABC and ∠ACB. Now, with the help of a compass and scale these angles are bisected.

AD, BE and CF are the three bisectors of the angles respectively. The bisectors of the angles meet at O. The point ‘O’ is called the in centre of the triangle.

Question 7. Let’s draw any angle ∠PQR using a scale, pencil and compass. Z PQR is bisected and the bisector of ∠RQP is produced up to point X. Again with scale, pencil and compass, ∠PQS is bisected by QY. What angle do QX and QY line segments make with each other? — Let us measure by protractor.
Solution:

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures Angle PQR Using Scale Pencil And Compass

Angle PQR is drawn using scale. Now, with help of compass ∠PQR is bisected. ∠PQR = 70°

QS is the bisector of ∠PQR.

∴ ∠RQS = ∠SQP = 35°.

Again, ∠PQS is bisected by QY with a compass.

∴ ∠SQY= ∠YQP = 17.5°.

∴ ∠XQY= ∠YQP = 17.5°.

Class 6 Math WBBSE Solutions

Question 8. Let’s draw a line segment PQ. At points P and Q of PQ, two perpendiculars PR and QS are drawn with scale, pencil and compass, ∠QPR and ∠PQS are bisected using scale, pencil and compass. The triangle so formed — its angles are measured and their values are written.
Solution:

WBBSE Solutions For Class 6 Maths Chapter Chapter 22 Drawing Of Different Geometrical Figures A Line Segments PQ At Point P And Q Of PQ

PQ is a line segment. At P and Q two perpendiculars RP and SQ are drawn with the help of scale and compass.

∴ ∠RPQ = ∠SQP = 90°.

Again, ∠RPQ and ∠SQP are bisected with the help of scale and compass. The bisectors meet each other at A. Therefore, ΔAPQ is formed.

Now, ∠APQ = 1/2 of 90° = 45°

and ∠AQP = 1/2 of 90° = 45°

∴ ∠PAQ = 180°- (45° + 45) = 180°- 90° = 90°.

WBBSE Solutions For Class 6 Maths Chapter 20 Geometrical Concept Of Circle

Class 6 Math Solutions WBBSE Chapter 20 Geometrical Concept Of Circle Exercise

Question 1. From the figure, let us answer the following :

WBBSE Solutions For Class 6 Maths Chapter 20 Geometrical Concept Of Circle O Is Center Of The Circle

1. O is the _____ of the circle.
Solution: O is the centre of the circle

2. Line segment OQ is the ____ of the circle.
Solution: radius of the circle.

3. Line segment PQ is the ______ of circle.
Solution: diameter of circle.

4. Line segment OP is the _______ of circle.
Solution: Line segment OP is radius of circle.

5. Line segment MN is ______ of the circle.
Solution: Line segment MN is a chord of the circle.

Read and Learn More WBBSE Solutions For Class 6 Maths

6. The points M and N have divided the circle into two _______ parts.
Solution: The points M and N have divided the circle into two arcs.

7. The area bounded by arc SR and radii SO and RO is called ______
Solution: The area bounded by the arc SR and radii SO and RO is called sector.

8. The diameter PQ has divided the circle into two equal parts which are called _____
Solution: The diameter PQ has divided the circle into two equal parts which are called semi-circles.

Class 6 Math Solutions WBBSE

Question 2. Let us write (✓) for the correct statement and (X) for the wrong one.

  1. All diameters of a circle are chords. [✓]
  2. All chords of a circle are its diameter. [X]
  3. The length of the diameter of a circle is twice the length of its radius. [✓]
  4. Sector of a circle is part of its circular area. [✓]
  5. Arc of a circle is part of the circle. [✓]
  6. Centre of the circle is a fixed point in the circle. [✓]
  7. Two diameters always intersect each of them. [✓]

Question 3. With the help of a compass and a pencil, a circle of radius 3 cm is drawn. Let’s name its centre, radius, diameter, chord, and arc.
Solution:

WBBSE Solutions For Class 6 Maths Chapter 20 Geometrical Concept Of Circle A Circle Of Radius 3 cm

  1. Centre O
  2. Radius OP
  3. Diameter AB
  4. Chord MN
  5. Arc PRB

Question 4. In this figure, let us colour the major segment yellow and minor segment green.

WBBSE Solutions For Class 6 Maths Chapter 20 Geometrical Concept Of Circle In Circel Major And Minor Segments
Solution:

  1. Major segment — yellow portion
  2. Minor segment — green portion

Class 6 Math Solutions WBBSE

Question 5. The radii of two circles are 2 cm and 4 cm respectively. Let’s write the lengths of their diameter without measuring.
Solution:

Radius of 1st circle = 2 cm.

Diameter of 1st circle = 2×2 cm. = 4 cm.

Radius of 2nd circle = 4 cm.

Diameter of 2nd circle = 2 x 4cm. = 8 cm.

Question 6. If the length of the greatest chord of a circle is 10 cm, let’s write what will be the length of its radius.
Solution:

Length of the greatest chord of a circle = 100 cm.

∴ Diameter of the circle = 10 cm.

∴ Radius of the circle = 10/2 cm = 5 cm.