You are on page 1of 1133

Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Heat and Thermodynamics
• Thermodynamics deals with heat, work,
temperature, and the statistical behavior of
systems with large numbers of particles
• Thermodynamics, broadly defined, is that
branch of physics that deals with the
relationships among macroscopic properties
of substances such as temperature, pressure,
volume and with the energy and flow of
energy associated with these properties.
Thermodynamics
• A knowledge of thermodynamics is essential
for understanding almost everything around
us, including our own body.
• Thermodynamics involves situations in which
the temperature or state (solid, liquid, gas) of
a system changes due to energy transfers.
Thermodynamics
• How the temperature of the light bulb in our
reading lamp is related to the amount of light
energy that it emits.
• How the ice cubes slowly melting in our
beverage keep the beverage cool.
• How the clothing we wear keeps us warm.
• How energy from food digestion keeps the
temperature of our body at a healthy value.
Temperature and the Zeroth
Law of Thermodynamics
• Temperature is a measure of the thermal
energy associated with the motion of particles
within a substance.
• We often associate the concept of
temperature with how hot or cold an object
feels when we touch it.
• Thus, our senses provide us with a qualitative
indication of temperature.
Temperature
• However, our senses are unreliable and often
mislead us.
• Example: Cardboard box + Metal tray from a
freezer
• Metal tray feels cooler even though the box is at
the same temperature
• Depends on the rate at which heat is transferred
from our hand to other objects.
• Scientists have developed instruments to
quantitatively measure the temperature.
Thermal Equilibrium
• Two objects with different temperature while in
contact eventually reach the same (some
intermediate) temperature.
• Example: Hot water + cold water = water with
intermediate temperature.
• Example: Two objects at different temperature in
contact with each other but isolated from the
environment.
• Since they are at different temperature, energy
will be exchanged.
• Two mechanisms: Heat exchange + E.M. Radiation
Thermal Contact &
Thermal Equilibrium
• Two objects are in thermal contact with each
other if energy can be exchanged between
them as either heat or radiation due to
temperature difference.
• Thermal equilibrium is a situation in which
two objects would not exchange energy by
heat or electromagnetic radiation if they were
placed in thermal contact.
Thermal Equilibrium
• Two objects: A and B not in thermal contact
• Third object: C -> our thermometer.

• When A and C are in thermal equilibrium, a


reading on C will be shown
• When B and C are in thermal equilibrium,
another (possibly) reading on C will be shown.
Zeroth Law of Thermodynamics
• If, however, the readings on the thermometer
are the same: object A and object B are in
thermal equilibrium
• If placed in thermal contact, there will be no
heat exchange.
• If objects A and B are separately in thermal
equilibrium with a third object C, then A and
B are in thermal equilibrium with each other.
Temperature
• Zeroth law of thermodynamics enables us to
define temperature:
• Temperature is the property that determines
whether an object is in thermal equilibrium with
another object.
• Two objects in thermal equilibrium with each
other are at the same temperature.
• Conversely, if two objects have different
temperatures, then they are not in thermal
equilibrium with each other.
Thermometers and the
Celsius Temperature Scale
• Thermometers are devices that measure temperature
of an object.
• In all thermometers, some physical property of a
system changes as its temperature changes.
• Examples:
(1) the volume of a liquid,
(2) the dimensions of a solid,
(3) The pressure of a gas at constant volume,
(4) the volume of a gas at constant pressure,
(5) the electric resistance of a conductor
(6) the color of an object, etc.
Thermometers and the
Celsius Temperature Scale
• Liquid—usually mercury or alcohol—that
expands into a glass capillary tube when
heated. Volume of the liquid changes.
• Temperature change in the range of the
thermometer is proportional to the length of
the liquid column.
• Thermometer can be calibrated by placing it in
thermal contact with some natural system
that remains at constant temperature.
Celsius Temperature Scale
• One reference system for calibration is a
mixture of water and ice in thermal
equilibrium at the atmospheric pressure.
• Celsius temperature scale:
• Zero degrees Celsius 0°C : ice point of water
(water and ice in thermal equilibrium at atmospheric pressure)
• Hundred degrees Celsius 100°C : steam point
of water
(water and steam in thermal equilibrium at atmospheric pressure)
Celsius Temperature Scale
• After setting reference points, the length of
the liquid column between the two points is
divided into 100 equal segments to create the
Celsius scale.
• Each segment denotes a change in
temperature of one Celsius degree.
Ideal Gas Temperature Scale
• Thermometers using liquid columns calibrated in
this way present problems when extremely
accurate readings are needed.
• Example: Alcohol thermometer calibrated at the
ice and steam points of water might agree with
those given by a mercury thermometer only at
the calibration points.
• Mercury and alcohol have different thermal
expansion properties,
• E.g. 50°C in mercury thermometer ≠ 50°C in
alcohol thermometer.
Ideal Gas Temperature Scale
• Discrepancies between thermometers are
especially large at temperatures measured far
from the calibration points.
• Additional practical problem: limited usable
range of temperatures.
• Example: freezing point of mercury, - 39°C,
boiling point of alcohol, 85°C
• We need a universal thermometer whose
readings are independent of the substance used
in it.
The Constant-Volume Gas
Thermometer
• Measures the pressure of the gas contained in
the flask immersed in the bath.
• The volume of gas in the flask
is kept constant.
• Raising or lowering the reservoir
B to keeps the mercury level
in column A constant.
The Constant-Volume Gas
Thermometer
• Physical change exploited is the variation of
pressure of a fixed volume of gas, with
respect to change in temperature.
• Initially calibrated by using the ice
and steam points of water.
• The flask was immersed in an ice-water bath and
mercury reservoir B was raised or lowered
• The top of the mercury in column A was at the
zero point on the scale.
The Constant-Volume Gas
Thermometer
• The height h, indicated the pressure in the flask.
• Flask was then immersed in water at the steam
point.
• Reservoir B was readjusted until the top of the
mercury in column A was again at the zero on the
scale.
• Ensured that the gas’s volume was the same as
that at the ice point temperature.
• This adjustment of reservoir B gave a value for
the gas pressure at 100°C.
The Constant-Volume Gas
Thermometer
• Two pressure and temperature values were then
plotted.
• The line connecting the two points serves as a
calibration curve for unknown temperatures.

𝑃𝑃 = ℎ𝜌𝜌𝜌𝜌 = ℎ 𝑇𝑇 𝜌𝜌𝜌𝜌

• To measure the temperature of a substance, place it in


thermal contact with the gas flask. Height h -> temperature
The Constant-Volume Gas
Thermometer
• Gas thermometers containing different gases at
different initial pressures.
• Experiments show that the thermometer
readings are nearly independent of the type of
gas used given:
(a) the gas pressure is low
(b) temperature is well above the liquefaction
temperature of the gases
• At lower pressures, agreement between
thermometers with different gases improves.
The Constant-Volume Gas
Thermometer

• If we extrapolate, we find a remarkable result:


In every case, the pressure is zero when the
temperature is -273.15°C !
Absolute Temperature Scale
• Suggests a special role played by -273.15 °C.
• Absolute temperature scale sets -273.15 °C as its
zero point – Absolute Zero Temperature
• Size of a degree on the absolute temperature
scale = size of a degree on the Celsius scale.
𝑇𝑇𝑐𝑐 = 𝑇𝑇 − 273.15
• Ice and steam points are experimentally difficult
to duplicate
• An absolute temperature scale based on two new
fixed points was adopted in 1954.
Absolute Temperature Scale
• The first reference point is absolute zero.
• The second reference temperature is the
triple point of water
• A single combination of temperature and
pressure at which liquid water, gaseous water,
and ice (solid water) coexist in equilibrium.
• Temperature of 0.01°C and a pressure of 4.58
mm of mercury.
Absolute Temperature Scale
• On the new i.e. Kelvin Scale, temperature of
water at the triple point = 273.16 K
• 1 K = 1/273.16 of the difference between
absolute zero and the temperature of the triple
point of water.
• What will happen at 0 K to gases if they do not
liquefy?
• Classically there will be no motion.
• Quantum mechanically there will still be some
zero-point-energy.
Celsius, Fahrenheit, and Kelvin Scales
• A common temperature scale in the US is
Fahrenheit scale: ice point 32°F steam point
212°F.
9
• 𝑇𝑇𝐹𝐹 = 𝑇𝑇𝐶𝐶 + 32𝑜𝑜 𝐹𝐹
5
5
• Δ𝑇𝑇𝐶𝐶 = Δ𝑇𝑇 = Δ𝑇𝑇𝐹𝐹
9
Macroscopic Description of an Ideal
Gas
• There is no equilibrium separation for the atoms
in a gas and hence there is no standard volume of
a gas at a given temperature.
• The equations involving gases will contain the
volume V as a variable.
• It is useful to know relationship between how the
quantities volume V, pressure P, and temperature
T for a sample of gas of mass m.
• Any equation relating these is called an equation
of state.
Ideal Gas
• In general, the equation of state of different
gases are different and quite complicated.
• However, if the gas is maintained at a very low
pressure (or low density), we get a simple and
universal equation of state.
• A real gas at very low pressure or density and
at a temperature well above the liquefaction
temperature of the gas is referred to as an
ideal gas.
Ideal Gas
• When the gas is kept at a constant
temperature, its pressure is inversely
proportional to its volume (Boyle’s law).
• When the pressure of the gas is kept constant,
its volume is directly proportional to its
temperature (the law of Charles and Gay-
Lussac).
• The equation of state for an ideal gas:
𝑃𝑃𝑃𝑃 = 𝑛𝑛𝑛𝑛𝑛𝑛
Ideal Gas
• Experimentally as the pressure approaches zero,
𝑃𝑃𝑃𝑃/( 𝑛𝑛𝑛𝑛) approaches the same value R for all
gases.
• R is called the universal gas constant (R = 8.314
J/mol-K = 0.082 14 L-atm/mol-K)
• One mole of any substance is that amount of the
substance that contains Avogadro’s number
𝑁𝑁𝐴𝐴 = 6.023 × 1023 of constituent particles.
𝑚𝑚
• Number of moles n of a substance : 𝑛𝑛 = ,
𝑀𝑀
M is the molar mass
Ideal Gas
• In terms of the total number of molecules N:
𝑁𝑁
• 𝑃𝑃𝑃𝑃 = 𝑛𝑛𝑛𝑛𝑛𝑛 = 𝑅𝑅𝑅𝑅 = 𝑁𝑁𝑘𝑘𝐵𝐵 𝑇𝑇
𝑁𝑁𝐴𝐴
𝑅𝑅
• 𝑘𝑘𝐵𝐵 = = 1.38 × 10−23 J/K
𝑁𝑁𝐴𝐴
• P, V, and T are the thermodynamic variables
of a gas.
Ideal gas
• Example-1: An ideal gas occupies a volume of
100 𝑐𝑐𝑚𝑚3 at 20°C and 100 Pa. Find the number
of moles of gas in the container.
• Solution: 𝑉𝑉 = 100 𝑐𝑐𝑚𝑚3 = 1.00 × 10−4 𝑚𝑚3 ,
𝑃𝑃 = 100 Pa, 𝑇𝑇 = 20𝑜𝑜 𝐶𝐶 = 293 𝐾𝐾
𝑃𝑃𝑃𝑃 100 𝑃𝑃𝑃𝑃 1.00 × 10−4 𝑚𝑚3
𝑛𝑛 = =
𝑅𝑅𝑅𝑅 𝐽𝐽
8.314 293 𝐾𝐾
𝑚𝑚𝑚𝑚𝑚𝑚. 𝐾𝐾
= 4.11 × 10−6 𝑚𝑚𝑚𝑚𝑚𝑚
Ideal Gas
• Filling a Scuba Tank: A certain scuba tank is
designed to hold 66.0 𝑓𝑓𝑡𝑡 3 of air when it is at
atmospheric pressure at 22°C. When this
volume of air is compressed to an absolute
pressure of 3 000 lb/𝑖𝑖𝑛𝑛2 and stored in a 10.0-L
(0.350- 𝑓𝑓𝑡𝑡 3 ) tank, the air becomes so hot that
the tank must be allowed to cool before it can
be used. Before the air cools, what is its
temperature? (Assume that the air behaves
like an ideal gas.)
Ideal Gas
• Solution: If no air escapes during the
compression, then the number of moles n of air
remains constant; therefore, using 𝑃𝑃𝑃𝑃 = 𝑛𝑛𝑛𝑛𝑛𝑛,
with n and R constant, we obtain a relationship
between the initial and final values:
𝑃𝑃𝑖𝑖 𝑉𝑉𝑖𝑖 𝑃𝑃𝑓𝑓 𝑉𝑉𝑓𝑓
• =
𝑇𝑇𝑖𝑖 𝑇𝑇𝑓𝑓
𝑃𝑃𝑓𝑓 𝑉𝑉𝑓𝑓 3000𝑙𝑙𝑙𝑙/𝑖𝑖𝑛𝑛2 0.350𝑓𝑓𝑡𝑡 3
• 𝑇𝑇𝑓𝑓 = 𝑇𝑇𝑖𝑖 = 295 K
𝑃𝑃𝑖𝑖 𝑉𝑉𝑖𝑖 14.7 𝑙𝑙𝑙𝑙/𝑖𝑖𝑛𝑛2 66.0𝑓𝑓𝑡𝑡 3
= 319 K
Heat and Internal Energy
• Zeroth law: If there is a temperature
difference there will be transfer of energy
between them.
• Heat: Heat is the energy that is transferred
across the boundary of a system due to a
temperature difference between the system
and its surroundings (i.e. between two
objects).
Heat
• When we heat a substance, we are
transferring energy into it by placing it in
contact with surroundings that have a higher
temperature.
• Example: pan of cold water on a stove burner.
• We also use the term heat to represent the
amount of energy transferred by this method
Internal Energy
• Internal Energy: Internal energy is all the energy of a
system that is associated with its microscopic
components—atoms and molecules—when viewed
from a reference frame at rest with respect to the
center of mass of the system.
• In the center of mass frame, there is no net motion of
the system.
1 1
• 𝑅𝑅𝑐𝑐𝑐𝑐 = ∑𝑚𝑚𝑖𝑖 𝑟𝑟𝑖𝑖 ⇒ 𝑉𝑉𝑐𝑐𝑐𝑐 = ∑𝑚𝑚𝑖𝑖 𝑣𝑣𝑖𝑖
M M
1
• ⇒0= ∑𝑚𝑚𝑖𝑖 𝑣𝑣𝑖𝑖 in the CoM system .
M
Internal Energy
• Bulk kinetic energy of the system due to its
motion through space is not included in internal
energy.
• Kinetic energy of random translational,
rotational, and vibrational motion of molecules
• potential energy within molecules
• potential energy between molecules
• Internal energy = thermal energy + potential
energy
Work and Mechanical Energy
• Work done on a system = amount of energy
transferred to the system from its surroundings.
• No sense of talking: “work of a system”
• We can only to the work done on or by a system.
• Mechanical energy of the system = kinetic and
potential energy of the system.
• Mechanical energy increases due to work done
on a system.
Heat and Work
• There is no sense in talking about the heat of
a system.
• Heat is the energy that has been transferred
as a result of a temperature difference.
• Although we refer to “latent heat” it is
actually “latent energy”.
• Both heat and work are ways of changing the
energy of a system.
Internal Energy and Work
• Internal energy can be changed by either heat
flow (to or from the system) or by work done
(by or on the system).
• Example-1: Gas in an insulated container is
compressed by a piston, the temperature of
the gas and its internal energy increase.
• No transfer of energy by heat from the
surroundings to the gas has occurred.
Units of Heat
• Calorie (cal) = the amount of energy transfer
necessary to raise the temperature of 1 g of
water from 14.5°C to 15.5°C.
• SI unit of heat energy= the joule.
• Mechanical Equivalent of Heat: 4.186 J of
mechanical energy raises the temperature of 1
g of water by 1°C from 14.5°C to 15.5°C.
• 1 cal ≡ 4.186 J mechanical equivalent of heat
Mechanical Equivalent of Heat
• Example-1: Losing Weight the Hard Way: A
student eats a dinner rated at 2 000 Calories.
He wishes to do an equivalent amount of work
in the gymnasium by lifting a 50.0-kg barbell.
How many times must he raise the barbell to
expend this much energy?
• (Assume that he raises the barbell 2.00 m
each time he lifts it and that he regains no
energy when he lowers the barbell.)
Mechanical Equivalent of Heat
• Solution: “Calorie,” written with a capital “C” and
used in describing the energy content of foods, is
actually a kilocalorie. 1 Calorie = 1 × 103 cal,
𝐽𝐽
• 𝑊𝑊 = 2.0 × 106 𝑐𝑐𝑐𝑐𝑐𝑐 4.186 = 8.37 × 106 𝐽𝐽
𝑐𝑐𝑐𝑐𝑐𝑐
• 𝑊𝑊 = 𝑛𝑛𝑛𝑛𝑛𝑛𝑛 = 8.37 × 106 𝐽𝐽
𝑊𝑊 8.37×106 𝐽𝐽
• 𝑛𝑛 = = = 8.54 × 103
𝑚𝑚𝑚𝑚𝑚 50.0 𝑘𝑘𝑘𝑘 ×(9.8 𝑚𝑚/𝑠𝑠 2 )(2.0 𝑚𝑚)
• If the student is in good shape and lifts the barbell once every
5 s, it will take him about 12 h to perform this feat.
• Some energy is used to pump blood, heat generation, etc.
Hence, in reality it will take less number of times to lift the
barbell.
Work and Heat in Thermodynamic
Processes
• State Variables: The state of a system in
thermodynamics is described by macroscopic
quantities or variables like pressure, volume,
temperature, internal energy.
• Transfer Variables: Variables in situations
involving energy transfer across a boundary.
• Transfer variables are zero unless a process occurs
in which energy is transferred across the
boundary of the system.
• Not associated with a given state of the system,
but with a change in the state of the system.
• Examples: Heat, work on or by the system.
Work on/by a System
• We consider work done on a deformable
system—a gas.
• A gas contained in a
cylinder fitted with a
movable piston.
• Work is done on a
gas as the piston is
pushed
Work on/by a System
• 𝑉𝑉 = Equilibrium pressure,
• 𝑃𝑃 = Uniform pressure on the cylinder’s wall.
• Force exerted by the gas on the piston
= 𝐹𝐹 = 𝑃𝑃𝑃𝑃, where 𝐴𝐴 = cross-sectional area
• We push the piston inward and compress the
gas very slowly such that the system remain in
thermal equilibrium at all times, with the
surroundings.
Quasi-static process
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Work and Heat in Thermodynamic
Processes
• State Variables: The state of a system in
thermodynamics is described by macroscopic
quantities or variables like pressure, volume,
temperature, internal energy.
• Transfer Variables: Variables in situations
involving energy transfer across a boundary.
• Transfer variables are zero unless a process occurs
in which energy is transferred across the
boundary of the system.
• Not associated with a given state of the system,
but with a change in the state of the system.
• Examples: Heat, work on or by the system.
Work on/by a System
• We consider work done on a deformable
system—a gas.
• A gas contained in a
cylinder fitted with a
movable piston.
• Work is done on a
gas as the piston is
pushed
Work on/by a System
• 𝑉𝑉 = Equilibrium pressure,
• 𝑃𝑃 = Uniform pressure on the cylinder’s wall.
• Force exerted by the gas on the piston
= 𝐹𝐹 = 𝑃𝑃𝑃𝑃, where 𝐴𝐴 = cross-sectional area
• We push the piston inward and compress the
gas very slowly such that the system remain in
thermal equilibrium at all times, with the
surroundings.
Quasi-static process
Work on/by a System
• 𝐹𝐹⃗ = −𝐹𝐹 𝚥𝚥̂ = External pushing force
• 𝑑𝑑 𝑟𝑟⃗ = 𝑑𝑑𝑑𝑑𝚥𝚥̂ = displacement of the piston
• Work done on the gas
• 𝑑𝑑𝑑𝑑 = 𝐹𝐹⃗ . 𝑑𝑑𝑟𝑟⃗ = −𝐹𝐹𝚥𝚥̂ . 𝑑𝑑𝑑𝑑𝚥𝚥̂ = −𝐹𝐹 𝑑𝑑𝑑𝑑 =
− 𝑃𝑃𝑃𝑃𝑃𝑃𝑃𝑃 = −𝑃𝑃𝑃𝑃𝑃𝑃
• The piston is always in equilibrium with the
external force and the force from the gas
molecules (Quasi-static process)
Work on/by a System
• Gas compressed:
𝒅𝒅𝒅𝒅 < 𝟎𝟎, 𝒅𝒅𝒅𝒅 = −𝑷𝑷 𝒅𝒅𝒅𝒅 > 𝟎𝟎, Work done on gas >0
• Gas expands:
𝒅𝒅𝒅𝒅 > 𝟎𝟎, 𝒅𝒅𝒅𝒅 = −𝑷𝑷 𝒅𝒅𝒅𝒅 < 𝟎𝟎, Work done on gas < 0
• Constant volume: Work done on the gas =0
• Total work done on the gas, for volume from 𝑉𝑉𝑖𝑖 to
𝑉𝑉𝑓𝑓 is
𝑉𝑉𝑓𝑓
𝑊𝑊 = − � 𝑃𝑃(𝑉𝑉) 𝑑𝑑𝑑𝑑
𝑉𝑉𝑖𝑖
PV- Diagram
• To evaluate the work done on
a system, we need 𝑃𝑃 = 𝑃𝑃(𝑉𝑉)
• If the pressure and volume are known at each step of
the process, the state of the gas at each step can be
plotted on a graph, called the PV diagram.
• Path between states: curve on a PV diagram
• The work done on a gas in a quasi-static process that
takes the gas from an initial state 𝑉𝑉𝑖𝑖 to a final state 𝑉𝑉𝑓𝑓
is the negative of the area under the curve on a PV
diagram, evaluated between the initial and final states.
PV- Diagram
• Work done depends on the particular path in
a PV diagram taken between the initial and
final states.

• The areas under the curves are different.


PV- Diagram
• The volume of the gas is first reduced from 𝑉𝑉𝑖𝑖
to 𝑉𝑉𝑓𝑓 at constant pressure 𝑃𝑃𝑖𝑖
• Then pressure of the gas then increases from
𝑃𝑃𝑖𝑖 to 𝑃𝑃𝑓𝑓 by heating at constant volume 𝑉𝑉𝑓𝑓
• Work done on the gas = −𝑃𝑃𝑖𝑖 (𝑉𝑉𝑓𝑓 − 𝑉𝑉𝑖𝑖 ) .
PV- Diagram
• Pressure of the gas is increased from 𝑃𝑃𝑖𝑖 to 𝑃𝑃𝑓𝑓
at constant volume 𝑉𝑉𝑖𝑖
• The volume of the gas is reduced from 𝑉𝑉𝑖𝑖 to 𝑉𝑉𝑓𝑓
at constant pressure 𝑃𝑃𝑓𝑓 .
• Work done on the gas = −𝑃𝑃𝑓𝑓 (𝑉𝑉𝑓𝑓 − 𝑉𝑉𝑖𝑖 ) .
Heat and Path
• Energy transfer 𝑄𝑄 into or out of a system by
heat also depends on the process.
• A gas at temperature 𝑇𝑇𝑖𝑖
expands slowly while
absorbing energy from
a reservoir in order to
maintain a constant
temperature.
• Energy reservoir is a source of energy that is considered to be so great that
a finite transfer of energy to or from the reservoir does not change its
temperature.
Heat and Path
• A gas expands rapidly into an
evacuated region after a
membrane is broken.
• Gas does no work because it
does not apply a force
• No force is required to expand into a vacuum
• No energy is transferred by heat through the
insulating wall.
Heat and Path
• The initial and final states of the ideal gas are
identical in two processes but paths are different.
• First process: Gas does work on the piston and
energy is transferred slowly to the gas by heat.
• Second process: No energy is transferred by heat,
and the value of the work done is zero.
• Energy transfer by heat, like work done,
depends on the initial, final, and intermediate
states of the system.
Heat, Work and Path

Heat and work depend on the


path, neither quantity is
determined solely by the end
points of a thermodynamic
process.
The First Law of Thermodynamics
• Two ways in which energy can be transferred
between a system and its surroundings:
(a) Work done on the system.
(b) Heat transferred to the system.
• The first law of thermodynamics is a special
case of the law of conservation of energy.
• It encompasses changes in internal energy and
energy transfer by heat and work.
The First Law of Thermodynamics
• Work results in a macroscopic displacement of
the point of application of a force.
• Heat transfer occurs on a molecular level
whenever a temperature difference exists
across the boundary of the system.
• Both mechanisms result in a change in the
internal energy of the system.
• Hence, result in measurable changes in the
macroscopic variables of the system (𝑃𝑃, 𝑉𝑉, 𝑇𝑇).
The First Law of Thermodynamics
• Consider a system undergoing a change from
an initial state 𝑉𝑉𝑖𝑖 , 𝑃𝑃𝑖𝑖 to a final state 𝑉𝑉𝑓𝑓 , 𝑃𝑃𝑓𝑓 .
During this:
• 𝑄𝑄 = Energy transfer by heat to the system
• 𝑊𝑊 = Work done on the system.
• The quantity 𝑄𝑄 + 𝑊𝑊 is measured for various
paths connecting the initial and final
equilibrium states.
• We find: 𝑸𝑸 + 𝑾𝑾 is independent of the path!
The First Law of Thermodynamics
• 𝑄𝑄 + 𝑊𝑊 depends solely on the initial and final
states.
• 𝑄𝑄 + 𝑊𝑊 must be function of the state, not of
the path.
• 𝑄𝑄 + 𝑊𝑊 must represent some change of
energy that is intrinsic to the state.
• 𝑄𝑄 + 𝑊𝑊 = Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = change in the internal
energy.
The First Law of Thermodynamics
• The change in internal energy of a system is equal
to the sum of the heat transferred to the system
and work done on the system.
Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄 + 𝑊𝑊
• In terms of infinitesimal quantities,
𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑑𝑑𝑄𝑄 + 𝑑𝑑𝑊𝑊
• First law of thermodynamics implies that there
exists a quantity known as internal energy whose
value is solely determined by the state of the
system.
The First Law of Thermodynamics
• Zeroth law of Thermodynamics : Concept of
temperature.
• First law of Thermodynamics: Concept of
internal energy.
• It is an energy conservation equation
specifying that the only type of energy that
changes in the system is the internal energy.
Applications of the First Law
• Example-1: Isolated System: one that does not
interact with its surroundings.
• No energy transfer by heat takes place and the
work done on the system is zero.
• Internal energy remains constant.
𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝑖𝑖 = 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝑓𝑓
• Internal energy 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 of an isolated system
remains constant.
Applications of the First Law
• Example-2: Cyclic Process: A system (one not
isolated from its surroundings) that is taken
through a cyclic process, i.e. process that
starts and ends at the same state.
• Change in the internal energy must again be
zero, since 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 is a state variable.
• Heat energy 𝑄𝑄 added to the system = negative
of the work W done on the system.
Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 0 ⇒ 𝑄𝑄 + 𝑊𝑊 = 0 ⇒ 𝑄𝑄 = −𝑊𝑊 (cyclic)
Applications of the First Law
• Cyclic Process (contd.): On a PV diagram, a cyclic
process appears as a closed curve.
• On the first path: 𝑑𝑑𝑑𝑑 = −𝑃𝑃𝑖𝑖 (𝑉𝑉𝑓𝑓 − 𝑉𝑉𝑖𝑖 )
• On the reverse of the second path: 𝑑𝑑𝑑𝑑 =
+ 𝑃𝑃𝑓𝑓 (𝑉𝑉𝑓𝑓 − 𝑉𝑉𝑖𝑖 )
• Total Work done
= (𝑃𝑃𝑓𝑓 − 𝑃𝑃𝑖𝑖 ) 𝑉𝑉𝑓𝑓 − 𝑉𝑉𝑖𝑖
=Area enclosed by the path
• In a cyclic process, the net work done on the system per cycle equals the area
enclosed by the path representing the process on a PV diagram.
Applications of the First Law
• Example-3: Adiabatic Process: Consider a sample of
gas contained in the piston–cylinder apparatus.
• An adiabatic process is one during which no energy
enters or leaves the system by heat —that is, 𝑄𝑄 = 0.
• An adiabatic process can be achieved
either by:
(a) Thermal insulation of the system
(b) Rapid process such that there is
insufficient time for energy transfer by
heat.
Applications of the First Law
• Adiabatic Process (contd.):
• Adiabatic process ≡ Process in a thermally insulated
system OR Adiabatic process ≡ Sudden Process.
• From the firs law: Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑊𝑊 (adiabatic process)
• Adiabatic compression: Work done on the system
𝑊𝑊 > 0 ⇒ Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 > 0, ⇒ Δ𝑇𝑇 > 0, temperature of the
system increases.
• Adiabatic expansion: Work done on the system
𝑊𝑊 < 0 ⇒ Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 < 0, ⇒ Δ𝑇𝑇 < 0, temperature of the
system decreases.
Applications of the First Law
• Adiabatic Process (contd.): Examples:
(a) expansion of hot gases in an internal combustion engine.
(b) liquefaction of gases in a cooling system
(thermally insulated).
(c) compression stroke in a diesel engine.
(d) Adiabatic Free Expansion: A gas expands
rapidly into an evacuated region in a
thermally insulated system after a
membrane is broken.
Q = W = 0 ⇒ Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 0.
• Initial and final internal energies of a gas are equal in an adiabatic
free expansion
Applications of the First Law
• Example-4: Isobaric Process: A process that
occurs at constant pressure.
• Put some fixed weight on the piston and allow
to move freely.
• 𝑃𝑃 = 𝑃𝑃𝑎𝑎𝑎𝑎𝑎𝑎 + 𝑀𝑀𝑀𝑀/𝐴𝐴 = 𝑃𝑃𝑔𝑔𝑔𝑔𝑔𝑔
• Piston is always at equilibrium with the net
force of gas pushing upward and force due to
weight applied and atmospheric pressure.
Applications of the First Law
• Isobaric Process (contd.): Both heat
transferred and work done are usually non-
zero.
• 𝑊𝑊 = −𝑃𝑃(𝑉𝑉𝑓𝑓 − 𝑉𝑉𝑖𝑖 ) Work done in isobaric process
• 𝑃𝑃 = constant pressure during the process.
Applications of the First Law
• Example-5: Isochoric Process: A process that
occurs at constant volume.
• In the cylinder-piston system, clamping the
piston at a fixed position would ensure an
isochoric (or isovolumetric) process.
• Work done is zero because the volume does
not change.
• Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄 + 0 = 𝑄𝑄 (isochoric process)
Applications of the First Law
• Isochoric Process (contd.): If energy is added by
heat to a system kept at constant volume, then all
of the transferred energy remains inside the
system resulting in an increase in its internal
energy.
• Example: Consider a can of spray paint is thrown
into fire. Energy enters into the system by heat
through the metal walls, the temperature, and
thus the pressure in the can increases and the
can eventually explodes.
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Applications of the First Law
• Example-1: Isolated System: one that does not
interact with its surroundings.
• No energy transfer by heat takes place and the
work done on the system is zero.
• Internal energy remains constant.
𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝑖𝑖 = 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝑓𝑓
• Internal energy 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 of an isolated system
remains constant.
Applications of the First Law
• Example-2: Cyclic Process: A system (one not
isolated from its surroundings) that is taken
through a cyclic process, i.e. process that
starts and ends at the same state.
• Change in the internal energy must again be
zero, since 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 is a state variable.
• Heat energy 𝑄𝑄 added to the system = negative
of the work W done on the system.
Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 0 ⇒ 𝑄𝑄 + 𝑊𝑊 = 0 ⇒ 𝑄𝑄 = −𝑊𝑊 (cyclic)
Applications of the First Law
• Cyclic Process (contd.): On a PV diagram, a cyclic
process appears as a closed curve.
• On the first path: 𝑑𝑑𝑑𝑑 = −𝑃𝑃𝑖𝑖 (𝑉𝑉𝑓𝑓 − 𝑉𝑉𝑖𝑖 )
• On the reverse of the second path: 𝑑𝑑𝑑𝑑 =
+ 𝑃𝑃𝑓𝑓 (𝑉𝑉𝑓𝑓 − 𝑉𝑉𝑖𝑖 )
• Total Work done
= (𝑃𝑃𝑓𝑓 − 𝑃𝑃𝑖𝑖 ) 𝑉𝑉𝑓𝑓 − 𝑉𝑉𝑖𝑖
=Area enclosed by the path
• In a cyclic process, the net work done on the system per cycle equals the area
enclosed by the path representing the process on a PV diagram.
Applications of the First Law
• Example-3: Adiabatic Process: Consider a sample of
gas contained in the piston–cylinder apparatus.
• An adiabatic process is one during which no energy
enters or leaves the system by heat —that is, 𝑄𝑄 = 0.
• An adiabatic process can be achieved
either by:
(a) Thermal insulation of the system
(b) Rapid process such that there is
insufficient time for energy transfer by
heat.
Applications of the First Law
• Adiabatic Process (contd.):
• Adiabatic process ≡ Process in a thermally insulated
system OR Adiabatic process ≡ Sudden Process.
• From the firs law: Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑊𝑊 (adiabatic process)
• Adiabatic compression: Work done on the system
𝑊𝑊 > 0 ⇒ Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 > 0, ⇒ Δ𝑇𝑇 > 0, temperature of the
system increases.
• Adiabatic expansion: Work done on the system
𝑊𝑊 < 0 ⇒ Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 < 0, ⇒ Δ𝑇𝑇 < 0, temperature of the
system decreases.
Applications of the First Law
• Adiabatic Process (contd.): Examples:
(a) expansion of hot gases in an internal combustion engine.
(b) liquefaction of gases in a cooling system
(thermally insulated).
(c) compression stroke in a diesel engine.
(d) Adiabatic Free Expansion: A gas expands
rapidly into an evacuated region in a
thermally insulated system after a
membrane is broken.
Q = W = 0 ⇒ Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 0.
• Initial and final internal energies of a gas are equal in an adiabatic
free expansion
Applications of the First Law
• Example-4: Isobaric Process: A process that
occurs at constant pressure.
• Put some fixed weight on the piston and allow
to move freely.
• 𝑃𝑃 = 𝑃𝑃𝑎𝑎𝑎𝑎𝑎𝑎 + 𝑀𝑀𝑀𝑀/𝐴𝐴 = 𝑃𝑃𝑔𝑔𝑔𝑔𝑔𝑔
• Piston is always at equilibrium with the net
force of gas pushing upward and force due to
weight applied and atmospheric pressure.
Applications of the First Law
• Isobaric Process (contd.): Both heat
transferred and work done are usually non-
zero.
• 𝑊𝑊 = −𝑃𝑃(𝑉𝑉𝑓𝑓 − 𝑉𝑉𝑖𝑖 ) Work done in isobaric process
• 𝑃𝑃 = constant pressure during the process.
Applications of the First Law
• Example-5: Isochoric Process: A process that
occurs at constant volume.
• In the cylinder-piston system, clamping the
piston at a fixed position would ensure an
isochoric (or isovolumetric) process.
• Work done is zero because the volume does
not change.
• Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄 + 0 = 𝑄𝑄 (isochoric process)
Applications of the First Law
• Isochoric Process (contd.): If energy is added by
heat to a system kept at constant volume, then all
of the transferred energy remains inside the
system resulting in an increase in its internal
energy.
• Example: Consider a can of spray paint is thrown
into fire. Energy enters into the system by heat
through the metal walls, the temperature, and
thus the pressure in the can increases and the
can eventually explodes.
Applications of the First Law
• Example-6: Isothermal Process: A process that
occurs at constant temperature.
• Example-a: In a cylinder-piston
system, putting the cylinder
in contact with some other
constant-temperature
reservoir makes the
process isothermal.
Applications of the First Law
• Isothermal Process (contd.):
• Example-b: Putting the cylinder inside an ice-
water bath.
• Example-c: When a substance freezes or boils,
the phase transition occurs at no change in
temperature.
• Isotherm: plot of P vs. V at constant temperature
for an ideal gas yields a rectangular hyperbolic
curve:
• 𝑃𝑃𝑃𝑃 = 𝑛𝑛𝑛𝑛𝑛𝑛 =const.
Applications of the First Law
• Isothermal Process (contd.): From first law:
Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 0 = 𝑄𝑄 + 𝑊𝑊 (Isothermal process)
𝑄𝑄 = −𝑊𝑊
• Energy transfer Q must be equal to the negative of the
work done on the gas.
• Any energy that enters the system by heat is
transferred out of the system by work.
• Any amount of work done on the system must be
transferred out of the system as heat.
• No change in the internal energy of the system occurs
in an isothermal process
Applications of the First Law
• Isothermal Process (contd.): For an ideal gas
undergoing an isothermal process, the quantity
of heat 𝑄𝑄 added to the gas is equal to the work
𝑊𝑊 done by the gas.
• Isothermal Expansion: 𝑄𝑄 > 0, heat flows into the
system ⇒ 𝑊𝑊 < 0 i.e. work on the system< 0 i.e.
system does work ⇒ Volume expands.
• Isothermal Compression: 𝑄𝑄 < 0, heat flows out of
the gas ⇒ 𝑊𝑊 > 0 i.e. work on by the system> 0
⇒ Volume decreases.
Applications of the First Law
• Isothermal Expansion of an Ideal Gas:
• An ideal gas is allowed to expand quasi-statically at
constant temperature (in contact with constant
temperature reservoir).
• Work done on the gas in changing the volume from 𝑉𝑉𝑖𝑖
to 𝑉𝑉𝑓𝑓 :
𝑉𝑉𝑓𝑓 𝑉𝑉𝑓𝑓
𝑉𝑉𝑓𝑓
𝑛𝑛𝑛𝑛𝑛𝑛 𝑑𝑑𝑑𝑑
𝑊𝑊 = − � 𝑃𝑃 𝑑𝑑𝑑𝑑 = − � 𝑑𝑑𝑑𝑑 = −𝑛𝑛𝑛𝑛𝑛𝑛 �
𝑉𝑉𝑖𝑖 𝑉𝑉 𝑉𝑉
𝑉𝑉𝑖𝑖 𝑉𝑉𝑖𝑖
𝑉𝑉𝑓𝑓
𝑊𝑊 = −𝑛𝑛𝑛𝑛𝑛𝑛 ln 𝑉𝑉 � = 𝑛𝑛𝑛𝑛𝑛𝑛ln 𝑉𝑉𝑖𝑖 /𝑉𝑉𝑓𝑓
𝑉𝑉𝑖𝑖
Applications of the First Law
• Isothermal Expansion of an Ideal Gas
(contd.): Work in an isothermal process = equals the
negative of the shaded area under the isotherm on a
PV curve.
Applications of the First Law
• Isothermal Expansion of an Ideal Gas
(contd.): Because the gas expands, 𝑉𝑉𝑓𝑓 >
𝑉𝑉𝑖𝑖 and the value for the work done on the gas
is negative.
• For isothermal compression of an ideal gas,
𝑉𝑉𝑓𝑓 < 𝑉𝑉𝑖𝑖 and the value for the work done on
the gas is positive.
Applications of the First Law
• Quiz-1: Fill in the boxes with +, - or 0. For each
situation, the system to be considered is
identified.
Situation System 𝑸𝑸 𝑾𝑾 𝚫𝚫𝐄𝐄𝐢𝐢𝐢𝐢𝐢𝐢
Rapidly pumping up a bicycle tire Air in the pump
Pan of room-temperature water sitting Water in the pan
on a hot stove
Air quickly leaking out of a balloon Air originally in the
balloon
Applications of the First Law
• Solution: (a) Because the pumping is rapid, no energy
enters or leaves the system by heat. Because 𝑊𝑊 > 0 when
work is done on the system, it is positive here. Thus, we see
that Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄 + 𝑊𝑊 must be positive. The air in the pump is
warmer.
Situation System 𝑸𝑸 𝑾𝑾 𝚫𝚫𝐄𝐄𝐢𝐢𝐢𝐢𝐢𝐢
Rapidly pumping up a bicycle tire Air in the pump 0 + +
Pan of room-temperature water sitting Water in the pan + 0 +
on a hot stove
Air quickly leaking out of a balloon Air originally in the 0 − −
balloon
Applications of the First Law
• Solution: (b) There is no work done either on or by the
system, but energy transfers into the water by heat from the
hot burner, making both 𝑄𝑄 and Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 positive.

Situation System 𝑸𝑸 𝑾𝑾 𝚫𝚫𝐄𝐄𝐢𝐢𝐢𝐢𝐢𝐢


Rapidly pumping up a bicycle tire Air in the pump 0 + +
Pan of room-temperature water sitting Water in the pan + 0 +
on a hot stove
Air quickly leaking out of a balloon Air originally in the 0 − −
balloon
Applications of the First Law
• Solution: (c) Again no energy transfers into or out of the
system by heat, but the air molecules escaping from the
balloon do work on the surrounding air molecules as they
push them out of the way. Thus 𝑊𝑊 is negative and Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 is
negative. The decrease in internal energy is evidenced by the
fact that the escaping air becomes cooler.
Situation System 𝑸𝑸 𝑾𝑾 𝚫𝚫𝐄𝐄𝐢𝐢𝐢𝐢𝐢𝐢
Rapidly pumping up a bicycle tire Air in the pump 0 + +
Pan of room-temperature water sitting Water in the pan + 0 +
on a hot stove
Air quickly leaking out of a balloon Air originally in the 0 − −
balloon
Applications of the First Law
• Quiz: Identify the nature of paths A, B, C, and
D (Q = 0 for path B):
Applications of the First Law
• Example-1: An Isothermal Expansion
• A 1.0-mol sample of an ideal gas is kept at 0.0°C
during an expansion from 3.0 L to 10.0 L.
(a) How much work is done on the gas during
the expansion?
(b) How much energy transfer by heat occurs
with the surroundings in this process?
(c) If the gas is returned to the original volume by
means of an isobaric process, how much work is
done on the gas?
Applications of the First Law
• Solution: The temperature of the ideal gas is
fixed. The process is isothermal.
• (a) Work done on the gas:
𝑊𝑊 = 𝑛𝑛𝑛𝑛𝑛𝑛 ln(𝑉𝑉𝑖𝑖 /𝑉𝑉𝑓𝑓 )
3.0L
= 1.0 mol 8.31 J/mol.K 273 K ln( )
10.0L
= −2.7 × 103 J
• (b) Energy is taken in by the gas as heat. From
first law:
• Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄 + 𝑊𝑊 = 0 ⇒ 𝑄𝑄 = −𝑊𝑊 = 2.7 × 103 J
Applications of the First Law
• Solution (contd.): (c) In an isobaric process, the
gas is returned to its original volume.
• We choose the path to consist of an isobaric process
from 𝑉𝑉𝑓𝑓 , 𝑃𝑃𝑓𝑓 → (𝑉𝑉𝑖𝑖 , 𝑃𝑃𝑓𝑓 ). In calculating 𝑃𝑃𝑓𝑓 , we use 𝑉𝑉𝑓𝑓
and 𝑇𝑇𝑖𝑖 because, we do not know the final temperature
𝑇𝑇 ′ after compression.
Applications of the First Law
• Solution (contd.): Initial volume= 10.0L=𝑉𝑉′𝑖𝑖 ,
final volume = 3.0L=𝑉𝑉′𝑓𝑓
𝑛𝑛𝑛𝑛𝑇𝑇𝑖𝑖′
• Pressure 𝑃𝑃 = 𝑃𝑃𝑓𝑓 = constant =
𝑉𝑉𝑖𝑖′
′ ′
• 𝑊𝑊 = −𝑃𝑃 𝑉𝑉𝑓𝑓 − 𝑉𝑉𝑖𝑖 =
1.0 𝑚𝑚𝑚𝑚𝑚𝑚 8.31 𝐽𝐽/𝑚𝑚𝑚𝑚𝑚𝑚.𝐾𝐾 273 𝐾𝐾
− 3.0 − 10.0 ×
10.0×10−3 𝑚𝑚3
−3 3
10 𝑚𝑚
= 1.6 × 103 𝐽𝐽
Applications of the First Law
• Example-2: Boiling Water Isobarically: 1.00 g of water
vaporizes isobarically at atmospheric pressure
(1.013 × 105 Pa). Its volume in the liquid state is
𝑉𝑉𝑖𝑖 = 𝑉𝑉𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙 = 1.00 𝑐𝑐𝑚𝑚3 , and its volume in the vapor
state is 𝑉𝑉𝑓𝑓 = 𝑉𝑉𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣 = 1671𝑐𝑐𝑚𝑚3 .
Find the work done in the expansion and the change
in internal energy of the system. (Ignore any mixing of
the steam and the surrounding air—imagine that the
steam simply pushes the surrounding air out of the
way.)
Applications of the First Law
• Solution: Because the expansion takes place at
constant pressure, the work done on the system
(the vaporizing water) as it pushes away the
surrounding air is:
𝑊𝑊 = −𝑃𝑃 𝑉𝑉𝑓𝑓 − 𝑉𝑉𝑖𝑖
= −(1.0 × 105 Pa)(1671.0−1.0) × 10−6 𝑚𝑚3
= −169 J
• The energy transfer 𝑄𝑄needed to vaporize water is
𝑄𝑄 = 𝑚𝑚𝐿𝐿𝑣𝑣 , where 𝑚𝑚 =mass of water, 𝐿𝐿𝑣𝑣 = latent
heat of vaporization for water = 2.26 × 106 J/kg
Applications of the First Law
• Solution (contd.):
• 𝑄𝑄 = 𝑚𝑚𝐿𝐿𝑣𝑣 = (1.0 × 10−3 kg)(2.26 × 106 )J/kg
• =2260 J
• Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄 + 𝑊𝑊 = 2260 J + −169J = 2.09 kJ
• The positive value for Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 indicates that the internal
energy of the system increases. Most of the energy
(2090 J)/(2260 J) = 93%, is transferred to the liquid
goes into increasing the internal energy of the system.
The remaining 7% of the energy transferred leaves the
system by work done by the steam on the surrounding
atmosphere.
Quiz
• 1. A sample of an ideal gas goes through the process shown in the
figure. From A to B, the process is adiabatic; from B to C, it is
isobaric with 100 kJ of energy entering the system by heat. From C
to D, the process is isothermal; from D to A, it is isobaric with 150 kJ
of energy leaving the system by heat.
• (A) Determine the work done on the system in path BC. [1]
• (B) Determine the work done on the system in path DA. [1]
• (C) Determine the difference in internal energy E int, B & E int, A. [3]
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Quiz
• 1. A sample of an ideal gas goes through the process shown in the
figure. From A to B, the process is adiabatic; from B to C, it is
isobaric with 100 kJ of energy entering the system by heat. From C
to D, the process is isothermal; from D to A, it is isobaric with 150 kJ
of energy leaving the system by heat.
• (A) Determine the work done on the system in path BC. [1]
• (B) Determine the work done on the system in path DA. [1]
• (C) Determine the difference in internal energy E int, B & E int, A. [3]
Quiz- Solution
• On BC: the process is
Isobaric.
1 atm=1.013 × 105 Pa
𝑊𝑊𝐵𝐵𝐵𝐵 = −𝑃𝑃(𝑉𝑉𝑓𝑓 − 𝑉𝑉𝑖𝑖 )
= −3.0atm 0.40𝑚𝑚3 − 0.09𝑚𝑚3 = −94.2kJ
• On DA: isobaric compression:
𝑊𝑊𝐷𝐷𝐷𝐷 = −1.0atm 0.2 − 1.2 𝑚𝑚3
= 1.013 × 105 Pa × 1.0𝑚𝑚3 = 101.3 kJ
Quiz- Solution
• On BC: Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐵𝐵𝐵𝐵 = 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐶𝐶 − 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐵𝐵 = 𝑄𝑄 + 𝑊𝑊
• Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐵𝐵𝐵𝐵 = 100 + −94.2 kJ = 5.8kJ
• On DA: Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐷𝐷𝐷𝐷 = 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐴𝐴 − 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐷𝐷 = 𝑄𝑄 + 𝑊𝑊
• Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐷𝐷𝐷𝐷 = −150 + 101.3 kJ = −48.7kJ
• On CD: Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 0, isothermal process.
⇒ 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐶𝐶 = 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐷𝐷
• Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐵𝐵𝐵𝐵 = 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐵𝐵 − 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐴𝐴 = 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐵𝐵 − 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐶𝐶 + 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐶𝐶 − 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐴𝐴 =
− Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐵𝐵𝐵𝐵 − Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐷𝐷𝐷𝐷 = −5.8 − −48.7 kJ
• 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐵𝐵 − 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐴𝐴 = 42.9 kJ Ans.
• HW: For a cyclic process: Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 0. Estimate the area enclosed
within the cycle and show that it is equal to the heat transferred by
the process.
Kinetic Theory of Gases
• Properties of an ideal gas can be described
using macroscopic variables as pressure,
volume, and temperature.
• Properties could also be described by using
the properties of the constituents of the gases
using averaging procedures.
• This is the approach of statistical mechanics.
Kinetic Theory of Gases
• Newton’s laws of motion applied in a
statistical manner to a collection of particles
provide a reasonable description of
thermodynamic processes.
• Because interaction between gas molecules is
much less than that of the liquids or solids,
statistical description is much easier.
Molecular Model of an Ideal Gas
• In Kinetic Theory, gas molecules move about in a
random fashion, colliding with the walls of their
container and with each other.
• Assumptions:
1. The number of molecules in the gas is large
~𝑁𝑁𝐴𝐴 .
2. Average separation between them is large
compared with their dimensions 𝜆𝜆 ≫ 𝑑𝑑, where
𝜆𝜆 is the mean free path, 𝑑𝑑 is molecular size. ⇒
molecules occupy a negligible volume in the
container.
Molecular Model of an Ideal Gas
3. Molecules move around randomly obeying
Newton’s laws of motion.
4. Molecules interact only by short-range forces
during elastic collisions.
5. Molecules make elastic collisions with the
walls.
6. The gas under consideration is a pure
substance i.e. all the molecules are identical.
Molecular Model of an Ideal Gas
• Behavior of molecular gases approximates
rather well that of ideal gases at low pressures
and well above the liquefaction temperature.
• Molecular rotations or vibrations have no
effect, on the average, on the motions.
• Ideally, we can approximate the molecules as
hard spheres, interacting only during
elastically collisions.
Kinetic Expression of Pressure
• Consider an ideal gas in a container of volume
𝑉𝑉 having 𝑁𝑁 molecules.
• 𝑑𝑑 =length of a side of cube.
• 𝑚𝑚 =mass of one molecule.
• 𝑣𝑣⃗𝑖𝑖 = velocity of i-th molecule
on which we focus our attention.
The molecule strikes a side wall making an
elastic collision and rebounds. (assumption-5)
Kinetic Expression of Pressure
• 𝑣𝑣𝑥𝑥𝑥𝑥 = 𝑥𝑥-component of the velocity
of the i-th molecule.
• Because of elastic collision, the
velocity component normal to
the wall is reversed, as mass of
the wall is much larger than that
of the molecule (wall does not
move). (assumption-3)
• 𝑦𝑦-component of velocity remains unchanged.
Kinetic Expression of Pressure
• Change in the 𝑥𝑥-component of the momentum
= Δ𝑝𝑝𝑥𝑥𝑥𝑥 = 𝑚𝑚𝑣𝑣𝑥𝑥𝑥𝑥−𝐹𝐹𝐹𝐹𝐹𝐹 − 𝑚𝑚𝑣𝑣𝑥𝑥𝑥𝑥−𝐼𝐼𝐼𝐼𝐼𝐼𝐼𝐼 = 𝑚𝑚 −𝑣𝑣𝑥𝑥𝑥𝑥 − 𝑚𝑚𝑣𝑣𝑥𝑥𝑥𝑥
• Δ𝑝𝑝𝑥𝑥𝑥𝑥 = −2𝑚𝑚𝑣𝑣𝑥𝑥𝑥𝑥
• From Newton’s 2nd law: Δ𝑝𝑝⃗ = ∫ 𝐹𝐹⃗ 𝑑𝑑𝑑𝑑,
• For a short impulsive collision, the force may be
approximated by an average force on the
molecule by the wall (assumption-3) :
• Δ𝑝𝑝𝑥𝑥𝑥𝑥 = 𝐹𝐹�𝑥𝑥𝑥𝑥,on molecule . Δ𝑡𝑡𝑐𝑐𝑐𝑐𝑐𝑐 = −2 𝑚𝑚𝑣𝑣𝑥𝑥𝑥𝑥
Kinetic Expression of Pressure
• Δ𝑡𝑡𝑐𝑐𝑐𝑐𝑐𝑐 = time of collision to the wall
• To a first approximation, the molecule travels
to the other side of the cube, moves back and
makes another collision with the first wall.
• Time interval between two collisions with the
same wall = Δ𝑡𝑡 = 2𝑑𝑑/𝑣𝑣𝑥𝑥𝑥𝑥
• Average force over Δ𝑡𝑡 time interval = Δ𝐹𝐹�𝑥𝑥𝑥𝑥 =
(change in the momentum during Δ𝑡𝑡) / Δ𝑡𝑡
Kinetic Expression of Pressure
• Average force on molecule between two collision
=
−2𝑚𝑚𝑣𝑣𝑥𝑥𝑥𝑥 2𝑚𝑚𝑣𝑣𝑥𝑥𝑥𝑥

Δ𝐹𝐹𝑥𝑥𝑖𝑖 = 2
= − 2𝑑𝑑 = −𝑚𝑚𝑣𝑣𝑥𝑥𝑥𝑥 /𝑑𝑑
Δ𝑡𝑡
𝑣𝑣𝑥𝑥𝑥𝑥

• From Newton’s third law, force on wall = −Δ𝐹𝐹�𝑖𝑖


2
𝑚𝑚𝑣𝑣𝑥𝑥𝑥𝑥
• 𝐹𝐹�𝑥𝑥𝑥𝑥,𝑜𝑜𝑜𝑜 𝑤𝑤𝑤𝑤𝑤𝑤𝑤𝑤 =
𝑑𝑑
• Total average force on wall due to all molecules =
𝑚𝑚 𝑁𝑁
� 𝑁𝑁 2 2
𝐹𝐹𝑥𝑥 = ∑𝑖𝑖=1 𝑚𝑚𝑣𝑣𝑥𝑥𝑥𝑥 /𝑑𝑑 = ∑𝑖𝑖=1 𝑣𝑣𝑥𝑥𝑥𝑥 (assumption-6)
𝑑𝑑
Kinetic Expression of Pressure
• For a small number of particles, the force on the
wall will vary with time as collision on wall will be
infrequent.
• For a large number of particles, variations in force
are smoothed out, so that the average force given
above is the same over any time interval.
(assumption-1)
• Constant force 𝐹𝐹𝑥𝑥 = 𝐹𝐹 on the wall due to the
𝑚𝑚 𝑁𝑁 2 𝑚𝑚
molecular collisions = 𝐹𝐹 = ∑𝑖𝑖=1 𝑣𝑣𝑥𝑥𝑥𝑥 = 𝑁𝑁𝑣𝑣𝑥𝑥2
𝑑𝑑 𝑑𝑑
Kinetic Expression of Pressure
1 𝑁𝑁 2
• 𝑣𝑣𝑥𝑥2
= ∑𝑖𝑖=1 𝑣𝑣𝑥𝑥𝑥𝑥
is the average squared 𝑥𝑥-component
𝑁𝑁
of velocity over the molecules.
• Because the motion is completely random,
𝑣𝑣𝑥𝑥2 = 𝑣𝑣𝑦𝑦2 = 𝑣𝑣𝑧𝑧2 for average over all the molecules.
(assumption-3)
• Hence from 𝑣𝑣 2 = 𝑣𝑣𝑥𝑥2 + 𝑣𝑣𝑦𝑦2 + 𝑣𝑣𝑧𝑧2 , we get
𝑁𝑁 𝑚𝑚𝑣𝑣 2 1 𝑁𝑁
• 𝐹𝐹 = = 𝑚𝑚 𝑣𝑣 2 𝑑𝑑2 = 𝑃𝑃 𝐴𝐴
3 𝑑𝑑 3 𝑑𝑑 3
𝐹𝐹 2 𝑁𝑁 1 2 𝑁𝑁
• 𝑃𝑃 = = 𝑚𝑚 𝑣𝑣 2 = 𝐾𝐾. 𝐸𝐸.
𝐴𝐴 3 𝑑𝑑 3 2 3 𝑉𝑉
Kinetic Expression of Pressure
• Pressure of a gas is proportional to the number of
molecules per unit volume and to the average
translational kinetic energy of the molecules.
• Relation of the macroscopic quantity of pressure
to a microscopic quantity— the average value of
the square of the molecular speed.
• Link between the molecular world and the large-
scale world
2 𝑁𝑁 1
𝑃𝑃 = 𝑚𝑚 𝑣𝑣 2
3 𝑉𝑉 2
Kinetic Expression of Pressure
• Ways to increase the pressure:
• (a) Increasing 𝑁𝑁/𝑉𝑉 i.e. the number of
molecules per unit volume
• (b) Increasing the average translational kinetic
energy of the gas molecules, which can be
done by increasing the temperature of the
gas.
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Quiz
• 1. A sample of an ideal gas goes through the process shown in the
figure. From A to B, the process is adiabatic; from B to C, it is
isobaric with 100 kJ of energy entering the system by heat. From C
to D, the process is isothermal; from D to A, it is isobaric with 150 kJ
of energy leaving the system by heat.
• (A) Determine the work done on the system in path BC. [1]
• (B) Determine the work done on the system in path DA. [1]
• (C) Determine the difference in internal energy E int, B & E int, A. [3]
Quiz- Solution
• On BC: the process is
Isobaric.
1 atm=1.013 × 105 Pa
𝑊𝑊𝐵𝐵𝐵𝐵 = −𝑃𝑃(𝑉𝑉𝑓𝑓 − 𝑉𝑉𝑖𝑖 )
= −3.0atm 0.40 − 0.09 𝑚𝑚3 = −94.2kJ
• On DA: isobaric compression:
𝑊𝑊𝐷𝐷𝐷𝐷 = −1.0atm 0.2 − 1.2 𝑚𝑚3
= 1.013 × 105 Pa × 1.0𝑚𝑚3 = 101.3 kJ
Quiz- Solution
• On BC: Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐵𝐵𝐵𝐵 = 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐶𝐶 − 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐵𝐵 = 𝑄𝑄 + 𝑊𝑊
• Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐵𝐵𝐵𝐵 = 100 + −94.2 kJ = 5.8kJ
• On DA: Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐷𝐷𝐷𝐷 = 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐴𝐴 − 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐷𝐷 = 𝑄𝑄 + 𝑊𝑊
• Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐷𝐷𝐷𝐷 = −150 + 101.3 kJ = −48.7kJ
• On CD: Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 0, isothermal process.
⇒ 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐶𝐶 = 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐷𝐷
• Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐵𝐵𝐵𝐵 = 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐵𝐵 − 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐴𝐴 = 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐵𝐵 − 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐶𝐶 + 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐶𝐶 −
𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐴𝐴 = −Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐵𝐵𝐵𝐵 − Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐷𝐷𝐷𝐷 = −5.8 − −48.7 kJ
• 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐵𝐵 − 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝐴𝐴 = 42.9 kJ Ans.
Quiz- Solution
• HW: For a cyclic process: Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 0. Estimate the area
enclosed within the cycle.
• Area inside the loop
= 2.0𝑎𝑎𝑎𝑎𝑎𝑎 × 0.4 − 0.09 𝑚𝑚3
3
1
− 0.2 − 0.09 𝑚𝑚 × × 2.0 𝑎𝑎𝑎𝑎𝑎𝑎
2
1.2 − 0.4 𝑚𝑚3
+2.0𝑎𝑎𝑎𝑎𝑎𝑎 ×
2
0.11 0.8
= 2.0𝑎𝑎𝑎𝑎𝑎𝑎 × 0.31 − + 𝑚𝑚3
2 2
= 0.655 × 2.0 × 1.013 × 105 𝑃𝑃𝑃𝑃 𝑚𝑚3 = 132.7𝑘𝑘𝑘𝑘
Kinetic Theory of Gases
• Properties of an ideal gas can be described
using macroscopic variables, such as pressure,
volume, and temperature.
• Properties could also be described by using
the properties of the constituents of the gases
using averaging procedures.
• This is the approach of statistical mechanics.
Kinetic Theory of Gases
• Newton’s laws of motion applied in a
statistical manner to a collection of particles
provide a reasonable description of
thermodynamic processes.
• Because interaction between gas molecules is
much less than that of the liquids or solids,
statistical description is much easier for a gas.
Molecular Model of an Ideal Gas
• In Kinetic Theory, gas molecules move about in a
random fashion, colliding with the walls of their
container and with each other.
• Assumptions:
1. The number of molecules in the gas is large
~𝑁𝑁𝐴𝐴 .
2. Average separation between them is large
compared with their dimensions 𝜆𝜆 ≫ 𝑑𝑑, where
𝜆𝜆 is the mean free path, 𝑑𝑑 is molecular size. ⇒
molecules occupy a negligible volume in the
container.
Molecular Model of an Ideal Gas
3. Molecules move around randomly obeying
Newton’s laws of motion.
4. Molecules interact only by short-range forces
during elastic collisions.
5. Molecules make elastic collisions with the
walls.
6. The gas under consideration is a pure
substance i.e. all the molecules are identical.
Molecular Model of an Ideal Gas
• Behavior of molecular gases approximates
rather well that of ideal gases at low pressures
and well above the liquefaction temperature.
• Molecular rotations or vibrations have no
effect, on the average, on the motions.
• Ideally, we can approximate the molecules as
hard spheres, interacting only during
elastically collisions.
Kinetic Expression of Pressure
• Consider an ideal gas in a container of volume
𝑉𝑉 having 𝑁𝑁 molecules.
• 𝑑𝑑 =length of a side of cube.
• 𝑚𝑚 =mass of one molecule.
• 𝑣𝑣⃗𝑖𝑖 = velocity of i-th molecule
on which we focus our attention.
The molecule strikes a side wall making an
elastic collision and rebounds. (assumption-5)
Kinetic Expression of Pressure
• 𝑣𝑣𝑥𝑥𝑥𝑥 = 𝑥𝑥-component of the velocity
of the i-th molecule.
• Because of elastic collision, the
velocity component normal to
the wall is reversed, as mass of
the wall is much larger than that
of the molecule (wall does not
move). (assumption-3)
• 𝑦𝑦-component of velocity remains unchanged.
Kinetic Expression of Pressure
• Change in the 𝑥𝑥-component of the momentum
= Δ𝑝𝑝𝑥𝑥𝑥𝑥 = 𝑚𝑚𝑣𝑣𝑥𝑥𝑥𝑥−𝐹𝐹𝐹𝐹𝐹𝐹 − 𝑚𝑚𝑣𝑣𝑥𝑥𝑥𝑥−𝐼𝐼𝐼𝐼𝐼𝐼𝐼𝐼 = 𝑚𝑚 −𝑣𝑣𝑥𝑥𝑥𝑥 − 𝑚𝑚𝑣𝑣𝑥𝑥𝑥𝑥
• Δ𝑝𝑝𝑥𝑥𝑥𝑥 = −2𝑚𝑚𝑣𝑣𝑥𝑥𝑥𝑥
• From Newton’s 2nd law: Δ𝑝𝑝⃗ = ∫ 𝐹𝐹⃗ 𝑑𝑑𝑑𝑑,
• For a short impulsive collision, the force may be
approximated by an average force on the
molecule by the wall (assumption-3) :
• Δ𝑝𝑝𝑥𝑥𝑥𝑥 = 𝐹𝐹�𝑥𝑥𝑥𝑥,on molecule . Δ𝑡𝑡𝑐𝑐𝑐𝑐𝑐𝑐 = −2 𝑚𝑚𝑣𝑣𝑥𝑥𝑥𝑥
Kinetic Expression of Pressure
• Δ𝑡𝑡𝑐𝑐𝑐𝑐𝑐𝑐 = time of collision to the wall
• To a first approximation, the molecule travels
to the other side of the cube, moves back and
makes another collision with the first wall.
• Time interval between two collisions with the
same wall = Δ𝑡𝑡 = 2𝑑𝑑/𝑣𝑣𝑥𝑥𝑥𝑥
• Average force over Δ𝑡𝑡 time interval = Δ𝐹𝐹�𝑥𝑥𝑥𝑥 =
(change in the momentum during Δ𝑡𝑡) / Δ𝑡𝑡
Kinetic Expression of Pressure
• Average force on molecule between two
collisions =
−2𝑚𝑚𝑣𝑣𝑥𝑥𝑥𝑥 2𝑚𝑚𝑣𝑣𝑥𝑥𝑥𝑥

Δ𝐹𝐹𝑥𝑥𝑖𝑖 = 2
= − 2𝑑𝑑 = −𝑚𝑚𝑣𝑣𝑥𝑥𝑥𝑥 /𝑑𝑑
Δ𝑡𝑡
𝑣𝑣𝑥𝑥𝑥𝑥

• From Newton’s third law, force on wall = −Δ𝐹𝐹�𝑖𝑖


2
𝑚𝑚𝑣𝑣𝑥𝑥𝑥𝑥
• 𝐹𝐹�𝑥𝑥𝑥𝑥,𝑜𝑜𝑜𝑜 𝑤𝑤𝑤𝑤𝑤𝑤𝑤𝑤 =
𝑑𝑑
• Total average force on wall due to all molecules =
𝑚𝑚 𝑁𝑁
� 𝑁𝑁 2 2
𝐹𝐹𝑥𝑥 = ∑𝑖𝑖=1 𝑚𝑚𝑣𝑣𝑥𝑥𝑥𝑥 /𝑑𝑑 = ∑𝑖𝑖=1 𝑣𝑣𝑥𝑥𝑥𝑥 (assumption-6)
𝑑𝑑
Kinetic Expression of Pressure
• For a small number of particles, the force on the
wall will vary with time as collision on wall will be
infrequent.
• For a large number of particles, variations in force
are smoothed out, so that the average force
given above is the same over any time interval.
(assumption-1)
• Constant force 𝐹𝐹𝑥𝑥 = 𝐹𝐹 on the wall due to the
𝑚𝑚 𝑁𝑁 2 𝑚𝑚
molecular collisions = 𝐹𝐹 = ∑𝑖𝑖=1 𝑣𝑣𝑥𝑥𝑥𝑥 = 𝑁𝑁𝑣𝑣𝑥𝑥2
𝑑𝑑 𝑑𝑑
Kinetic Expression of Pressure
1 𝑁𝑁 2
• 𝑣𝑣𝑥𝑥2
= ∑𝑖𝑖=1 𝑣𝑣𝑥𝑥𝑥𝑥
is the average squared 𝑥𝑥-component
𝑁𝑁
of velocity over the molecules.
• Because the motion is completely random,
𝑣𝑣𝑥𝑥2 = 𝑣𝑣𝑦𝑦2 = 𝑣𝑣𝑧𝑧2 for average over all the molecules.
(assumption-3)
• Hence from 𝑣𝑣 2 = 𝑣𝑣𝑥𝑥2 + 𝑣𝑣𝑦𝑦2 + 𝑣𝑣𝑧𝑧2 , we get
𝑁𝑁 𝑚𝑚𝑣𝑣 2 1 𝑁𝑁
• 𝐹𝐹 = = 𝑚𝑚 𝑣𝑣 2 𝑑𝑑2 = 𝑃𝑃 𝐴𝐴
3 𝑑𝑑 3 𝑑𝑑 3
𝐹𝐹 2 𝑁𝑁 1 2 𝑁𝑁
• 𝑃𝑃 = = 𝑚𝑚 𝑣𝑣 2 = 𝐾𝐾. 𝐸𝐸.
𝐴𝐴 3 𝑑𝑑 3 2 3 𝑉𝑉
Kinetic Expression of Pressure
• Pressure of a gas is proportional to the number of
molecules per unit volume and to the average
translational kinetic energy of the molecules.
• A relation of the macroscopic quantity of
pressure to a microscopic quantity— the average
value of the square of the molecular speed.
• Link between the molecular world and the large-
scale world
2 𝑁𝑁 1
𝑃𝑃 = 𝑚𝑚 𝑣𝑣 2
3 𝑉𝑉 2
Kinetic Expression of Pressure
• Ways to increase the pressure:
• (a) Increasing 𝑁𝑁/𝑉𝑉 i.e. the number of
molecules per unit volume
• (b) Increasing the average translational kinetic
energy of the gas molecules, which can be
done by increasing the temperature of the
gas.
Molecular Interpretation of
Temperature
• For an ideal gas from Kinetic Theory:
2 𝑁𝑁 1 2 1
• 𝑃𝑃 = 𝑚𝑚 𝑣𝑣 2 ⇒ 𝑃𝑃𝑃𝑃 = 𝑁𝑁 𝑚𝑚 𝑣𝑣 2
3 𝑉𝑉 2 3 2
• From equation of state: 𝑃𝑃𝑃𝑃 = 𝑁𝑁𝑘𝑘𝐵𝐵 𝑇𝑇
2 1
• 𝑇𝑇 = ( 𝑚𝑚𝑣𝑣 2 )
3𝑘𝑘𝐵𝐵 2
• Temperature is a direct measure of average
molecular kinetic energy.
1 3
• 𝑚𝑚𝑣𝑣 2 = 𝑘𝑘𝐵𝐵 𝑇𝑇
2 2
Translational Kinetic Energy
1 3
• 𝐾𝐾. 𝐸𝐸. = 𝑚𝑚𝑣𝑣 2
= average over the
𝑘𝑘𝐵𝐵 𝑇𝑇
2 2
molecules
• Total molecular kinetic energy of 𝑁𝑁 molecules
𝑁𝑁 3𝑁𝑁 3
𝑚𝑚𝑣𝑣 2
• 𝐾𝐾𝑡𝑡𝑡𝑡𝑡𝑡,𝑡𝑡𝑡𝑡𝑡𝑡 = 𝑁𝑁𝐾𝐾. 𝐸𝐸. = = = 𝑘𝑘𝐵𝐵 𝑇𝑇 𝑛𝑛𝑛𝑛𝑛𝑛
2 2 2
• For gas in which molecules possess only
translational kinetic energy, the above is the
internal energy of the gas.
• Internal energy of an ideal gas depends only on
temperature.
Equipartition of Energy
• For random motion of the gas molecules,
1 2
• 𝑣𝑣𝑥𝑥2
= 𝑣𝑣𝑦𝑦2 = 𝑣𝑣𝑧𝑧2 ⇒ 𝑣𝑣𝑥𝑥2 = 2 2
𝑣𝑣𝑦𝑦 = 𝑣𝑣𝑧𝑧 = 𝑣𝑣
3
1 2 1 1 2 1
• Hence, 𝑚𝑚𝑣𝑣𝑥𝑥 = 𝑘𝑘𝐵𝐵 𝑇𝑇 = 𝑚𝑚𝑣𝑣𝑦𝑦 = 𝑚𝑚𝑣𝑣𝑧𝑧2
2 2 2 2
• Thus, each translational degree of freedom
contributes an equal amount of energy.
• A degree of freedom refers to an independent
means by which a molecule can possess
energy.
Equipartition of Energy
• Theorem of Equipartition of Energy states:
Each degree of freedom in an ideal gas
contributes equal amount of energy to the
total energy of a system.
• For an ideal gas, each translational degree of
1
freedom contributes 𝑘𝑘𝐵𝐵 𝑇𝑇 to the gas.
2
RMS (Root-Mean-Square) Speed
• A measure of how fast the gas molecules
move.
1 𝑁𝑁
• 𝑣𝑣𝑟𝑟𝑟𝑟𝑟𝑟 = 𝑣𝑣 2 = ∑𝑖𝑖=1 𝑣𝑣𝑖𝑖2
𝑁𝑁

• 𝑣𝑣𝑟𝑟𝑟𝑟𝑟𝑟 = 3 𝑘𝑘𝐵𝐵 𝑇𝑇/𝑚𝑚 = 3𝑅𝑅𝑅𝑅/𝑀𝑀


• 𝑀𝑀 =molar mass in kilograms per mole=𝑚𝑚𝑁𝑁𝐴𝐴
• Lighter molecules move faster, on an average,
than heavier molecules (at fixed 𝑇𝑇).
RMS (Root-Mean-Square) Speed
Gas Molar mass 𝒗𝒗𝒓𝒓𝒓𝒓𝒓𝒓 at 𝟐𝟐𝟎𝟎𝒐𝒐 C(m/s)
(g/mol)
𝐻𝐻2 2.02 1902
𝐻𝐻𝐻𝐻 4.00 1352
𝐻𝐻2 𝑂𝑂 18.0 637
𝑁𝑁𝑁𝑁 20.2 602
𝑁𝑁2 or 𝐶𝐶𝐶𝐶 28.0 511
𝑁𝑁𝑁𝑁 30.0 494
𝑂𝑂2 32.0 478
𝐶𝐶𝑂𝑂2 44.0 408
Kinetic Expression of Pressure
• Quiz-A: Two containers hold an ideal gas at the
same temperature and pressure. Both containers
hold the same type of gas but container B has
twice the volume of container A.
• The average translational kinetic energy per
molecule in container B is
(a) twice that for container A
(b) the same as that for container A
(c) half that for container A
(d) impossible to determine.
Kinetic Expression of Pressure
• Quiz-B: The internal energy of the gas in
container B is
(a) twice that for container A
(b) the same as that for container A
(c) half that for container A
(d) impossible to determine.
Kinetic Expression of Pressure
• Quiz-C: The rms speed of the gas molecules in
container B is
(a) twice that for container A
(b) the same as that for container A
(c) half that for container A
(d) impossible to determine.
Kinetic Expression of Pressure
• Answers:
• A- The average translational kinetic energy per
molecule is a function only of temperature.
• B- Because there are twice as many molecules
and the temperature of both containers is the
same, the total energy in B is twice that in A.
• C- Because both containers hold the same
type of gas, the rms speed is a function only of
temperature.
Kinetic Interpretation of Temperature
• Example: A Tank of Helium
• A tank used for filling helium balloons has a
volume of 0.300 𝑚𝑚3 and contains 2.00 mol of
helium gas at 20.0°C. Assume that the helium
behaves like an ideal gas.
• (A) What is the total translational kinetic
energy of the gas molecules?
• (B) What is the average kinetic energy per
molecule?
Kinetic Interpretation of Temperature
3 3
• Solution: (A) 𝐾𝐾𝑡𝑡𝑡𝑡𝑡𝑡,𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡 = 𝑁𝑁𝑘𝑘𝐵𝐵 𝑇𝑇 = 𝑛𝑛𝑛𝑛𝑛𝑛
2 2
3
• 𝐾𝐾𝑡𝑡𝑡𝑡𝑡𝑡,𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡 = (2.00 𝑚𝑚𝑚𝑚𝑚𝑚)(8.31𝐽𝐽/𝑚𝑚𝑚𝑚𝑚𝑚. 𝐾𝐾)(293𝐾𝐾)
2
• = 7.30 × 103 𝐽𝐽
1 3 3 10−23 𝐽𝐽
• (B) 𝑚𝑚 𝑣𝑣 2 = 𝑘𝑘𝐵𝐵 𝑇𝑇
= 1.38 × 293 𝐾𝐾 =
2 2 2 𝐾𝐾
−21 𝐽𝐽
6.07 × 10
• What if the temperature is raised from 20.0°C to
40.0°C? Because 40.0 is twice as large as 20.0, is the
total translational energy of the molecules of the gas
twice as large at the higher temperature?
Molar Specific Heats of an Ideal Gas
• Consider an ideal gas undergoing several
processes such that change in temperature is
the same: Δ𝑇𝑇 = 𝑇𝑇𝑓𝑓 − 𝑇𝑇𝑖𝑖
• Temperature can change
in different paths.
• Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 same for all paths
• 𝑄𝑄 = Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 − 𝑊𝑊 different
• 𝑊𝑊 is different (Why?)
Molar Specific Heats of an Ideal Gas
Heat associated with a given change in
temperature does not have a unique value.
• Number of moles is a convenient measure of the
amount of gas. Hence, we define molar specific
heats.
𝑄𝑄 = 𝑛𝑛𝐶𝐶𝑉𝑉 Δ𝑇𝑇 (constant volume)
𝑄𝑄 = 𝑛𝑛𝐶𝐶𝑃𝑃 Δ𝑇𝑇 (constant pressure)
• Molar specific heat at constant volume 𝐶𝐶𝑉𝑉
• Molar specific heat at constant pressure 𝐶𝐶𝑃𝑃
Molar Specific Heats of an Ideal Gas
• Constant Pressure Specific Heat: Add energy to a
gas by heat at constant pressure => 𝑄𝑄 > 0
• Internal energy of the gas increase, Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 > 0
• Work is done by the gas because of the change in
volume ( Δ𝑉𝑉 > 0), ⇒ 𝑊𝑊𝑜𝑜𝑜𝑜 = 𝑊𝑊 < 0
• 𝑄𝑄𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐 accounts for both the increase in internal
energy and the transfer of energy out of the
system by work:
𝑄𝑄𝑃𝑃 = 𝑛𝑛𝐶𝐶𝑃𝑃 Δ𝑇𝑇 = Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 − 𝑊𝑊
Molar Specific Heats of an Ideal Gas
• Constant Volume Specific Heat: Add energy to
a gas by heat at constant volume => 𝑄𝑄 > 0
• At constant volume, no work is done on/by
the gas: 𝑊𝑊 = −∫ 𝑃𝑃𝑃𝑃𝑃𝑃 = 0, since 𝑉𝑉 = 𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐
• Hence, 𝑄𝑄𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐 accounts for only increase of
internal energy.
• At constant volume, Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄 = 𝑛𝑛𝐶𝐶𝑉𝑉 Δ𝑇𝑇
• ⇒ 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑛𝑛𝐶𝐶𝑉𝑉 𝑇𝑇 Internal energy of an ideal gas
Molar Specific Heats of an Ideal Gas
• 𝑖𝑖 → 𝑓𝑓: isochoric process
• All heat goes into increasing
internal energy. No work is
done on/by system.
• 𝑖𝑖 → 𝑓𝑓𝑓: isobaric process
• Part of the energy transferred in by heat is
used (a) to increase 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖
and part is (b) transferred out by work.
Molar Specific Heat of an Ideal Gas
• For a mono-atomic gas, Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝐾𝐾𝑡𝑡𝑡𝑡𝑡𝑡,𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡 =
3
𝑛𝑛𝑛𝑛𝑛𝑛
2
• At constant volume, 𝑊𝑊 = 0, 𝑄𝑄 = Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 =
𝑛𝑛𝐶𝐶𝑉𝑉 Δ𝑇𝑇
• For constant molar specific
heat (𝐶𝐶𝑉𝑉 ): 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑛𝑛𝐶𝐶𝑉𝑉 𝑇𝑇
𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 3 3
• 𝐶𝐶𝑉𝑉 = = 𝑛𝑛𝑛𝑛𝑛𝑛/(𝑛𝑛𝑛𝑛) = 𝑅𝑅
𝑛𝑛𝑛𝑛 2 2
Molar Specific Heat at Constant
Volume of an Ideal Gas
1 𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖
• Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑛𝑛𝐶𝐶𝑉𝑉 Δ𝑇𝑇 ⇒ 𝐶𝐶𝑉𝑉 =
𝑛𝑛 𝑑𝑑𝑑𝑑
in infinitesimal
form.
3
• 𝐶𝐶𝑉𝑉 = 𝑅𝑅 = 12.5 J/mol.K for all monoatomic gas.
2
Molar Specific Heats for monoatomic Gases
Molar Specific Heat J/mol.K (at 300K)
Gas 𝐶𝐶𝑃𝑃 𝐶𝐶𝑉𝑉 𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉 𝛾𝛾 = 𝐶𝐶𝑃𝑃 /𝐶𝐶𝑉𝑉
He 20.8 12.5 8.33 1.67
Ar 20.8 12.5 8.33 1.67
Ne 20.8 12.7 8.12 1.64
Kr 20.8 12.3 8.49 1.69
Relation Between Molar Specific
Heats of an Ideal Gas
• For a process at constant volume, for an ideal
gas, Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄𝑉𝑉 = 𝑛𝑛𝐶𝐶𝑉𝑉 Δ𝑇𝑇
• For a process at constant pressure, for an ideal
gas: 𝑄𝑄𝑃𝑃 = 𝑛𝑛𝐶𝐶𝑃𝑃 Δ𝑇𝑇 = Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 − 𝑊𝑊
• Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄𝑃𝑃 + 𝑊𝑊 = 𝑛𝑛𝐶𝐶𝑃𝑃 Δ𝑇𝑇 + −𝑃𝑃Δ𝑉𝑉
• ⇒ 𝑛𝑛𝐶𝐶𝑉𝑉 Δ𝑇𝑇 = 𝑛𝑛𝐶𝐶𝑃𝑃 Δ𝑇𝑇 − 𝑛𝑛𝑛𝑛Δ𝑇𝑇 Δ 𝑃𝑃𝑃𝑃 = 𝑃𝑃Δ𝑉𝑉
⇒ 𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉 = 𝑅𝑅 for an ideal gas
Relation Between Molar Specific
Heats of an Ideal Gas
3
• For monatomic gas: 𝐶𝐶𝑉𝑉 = 𝑅𝑅, 𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉 = 𝑅𝑅
2
3 3 5
• ⇒ 𝐶𝐶𝑃𝑃 /𝐶𝐶𝑉𝑉 = ( + 1)/( ) = = 1.67
2 2 3
• For monoatomic gas, experimental value of 𝛾𝛾
is very close to theoretical value.
• For diatomic or triatomic gas, the
experimental value of 𝛾𝛾 is quite different from
1.67.
Relation Between Molar Specific
Heats of an Ideal Gas
• Some additional contribution to the molar
specific heat occurs from the internal
structure of the more complex molecules.
• The internal energy—and, hence, the molar
specific heat—of a complex gas must include
contributions from the rotational and the
vibrational motions of the molecule.
Adiabatic Processes for an Ideal Gas
• No energy is transferred by heat between a
system and its surroundings.
• Either a process in thermal insulation or a
sudden process.
• Example: A gas is compressed (or expanded)
very rapidly, very little energy is transferred
out of (or into) the system by heat => nearly
adiabatic.
• E.g. the cycle of a gasoline engine.
Adiabatic Processes for an Ideal Gas
• Consider a gas undergoing adiabatic
expansion/compression quasi-statically such
that:
• (a) The system is at a thermodynamic
equilibrium state at any particular moment.
• (b) The ideal gas equation of state 𝑃𝑃𝑃𝑃 =
𝑛𝑛𝑛𝑛𝑛𝑛remains valid.
• In an adiabatic process, all three
thermodynamic variables 𝑃𝑃, 𝑉𝑉 and 𝑇𝑇 changes.
Adiabatic Processes for an Ideal Gas
• Adiabatic Compression: Suppose that an ideal
gas is compressed adiabatically quasi-statically.
• Heat transferred 𝑄𝑄 = 0.
• 𝑑𝑑𝑑𝑑 = infinitesimal change in volume
• 𝑑𝑑𝑑𝑑 = Work done on the gas = −𝑃𝑃𝑃𝑃𝑃𝑃
• 𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = infinitesimal change in internal
energy (𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 ≠ 0, since 𝑑𝑑𝑑𝑑 ≠ 0)
• 𝑑𝑑𝑑𝑑 = infinitesimal change in temperature of
gas ( 𝑑𝑑𝑑𝑑 ≠ 0, since 𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 ≠ 0)
Adiabatic Processes for an Ideal Gas
• Adiabatic Compression (contd.): But we know for
an ideal gas: 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑛𝑛 𝐶𝐶𝑉𝑉 𝑇𝑇 ⇒ 𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑛𝑛𝐶𝐶𝑉𝑉 𝑑𝑑𝑑𝑑
• From first law: 𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑 = −𝑃𝑃𝑃𝑃𝑃𝑃 ⇒
𝑛𝑛𝐶𝐶𝑉𝑉 𝑑𝑑𝑑𝑑 = −𝑃𝑃𝑃𝑃𝑃𝑃 (adiabatic process)
• But 𝑃𝑃𝑃𝑃 = 𝑛𝑛𝑛𝑛𝑛𝑛 ⇒ 𝑃𝑃𝑃𝑃𝑃𝑃 + 𝑑𝑑𝑑𝑑 𝑉𝑉 = 𝑛𝑛𝑛𝑛𝑛𝑛𝑛𝑛
𝑛𝑛𝑛𝑛𝑛𝑛
• 𝑃𝑃𝑃𝑃𝑃𝑃 + 𝑉𝑉𝑉𝑉𝑉𝑉 = − 𝑑𝑑𝑑𝑑 , eliminating 𝑑𝑑𝑑𝑑
𝑛𝑛𝐶𝐶𝑉𝑉
• Using 𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉 = 𝑅𝑅, we get 𝑃𝑃𝑃𝑃𝑃𝑃 + 𝑉𝑉𝑉𝑉𝑉𝑉 =
𝐶𝐶𝑃𝑃
1− 𝑃𝑃𝑃𝑃𝑃𝑃 = 1 − 𝛾𝛾 𝑃𝑃𝑃𝑃𝑃𝑃
𝐶𝐶𝑉𝑉
Adiabatic Processes for an Ideal Gas
• Adiabatic Compression (contd.): Hence,
𝑑𝑑𝑑𝑑 𝑑𝑑𝑑𝑑
𝑉𝑉𝑉𝑉𝑉𝑉 = −𝛾𝛾𝛾𝛾 𝑑𝑑𝑑𝑑 ⇒ + 𝛾𝛾 = 0
𝑃𝑃 𝑉𝑉
• Integrating, we get: ln 𝑃𝑃 + 𝛾𝛾 ln 𝑉𝑉 = 𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐.
• Adiabatic process: 𝑃𝑃𝑉𝑉 𝛾𝛾 = 𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐.
• The same is true for adiabatic expansion
(𝑑𝑑𝑑𝑑 < 0, 𝑑𝑑𝑑𝑑 > 0, 𝑑𝑑𝑑𝑑 < 0 )
𝛾𝛾 𝛾𝛾
𝑃𝑃𝑖𝑖 𝑉𝑉𝑖𝑖 = 𝑃𝑃𝑓𝑓 𝑉𝑉𝑓𝑓 = 𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐.
Adiabatic Processes for an Ideal Gas
• Adiabatic Compression: 𝑑𝑑𝑑𝑑 = 0, 𝑑𝑑𝑑𝑑 > 0,
internal energy increases 𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 > 0 ⇒ 𝑑𝑑𝑑𝑑 > 0.
• Temperature of a gas increases in adiabatic
compression.
• Adiabatic Expansion : 𝑑𝑑𝑑𝑑 = 0, 𝑑𝑑𝑑𝑑 < 0, internal
energy decreases 𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 < 0 ⇒ 𝑑𝑑𝑑𝑑 < 0.
• Temperature of a gas decreases in adiabatic
expansion. (compare with adiabatic free
expansion in which 𝑇𝑇 = 𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐. Why? )
Adiabatic Processes for an Ideal Gas
• Adiabat: In a 𝑃𝑃𝑃𝑃-diagram, the path of an
adiabatic process is called an adiabat.
𝛾𝛾 𝐶𝐶𝑃𝑃
• Since 𝑃𝑃𝑉𝑉 = 𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐. and 𝛾𝛾 = > 1, adiabat is
𝐶𝐶𝑉𝑉
steeper than an isotherm (𝑃𝑃𝑃𝑃 = 𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐.)
• Adiabatic process in ideal gas:
• 𝑃𝑃𝑉𝑉 𝛾𝛾 = 𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐. and 𝑃𝑃𝑃𝑃 = 𝑛𝑛𝑛𝑛𝑛𝑛 ≠ 𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐.
𝛾𝛾 𝛾𝛾 𝛾𝛾−1 𝛾𝛾−1
• 𝑃𝑃𝑖𝑖 𝑉𝑉𝑖𝑖 = 𝑃𝑃𝑓𝑓 𝑉𝑉𝑓𝑓 ⇒ 𝑃𝑃𝑖𝑖 𝑉𝑉𝑖𝑖 𝑉𝑉𝑖𝑖 = 𝑃𝑃𝑓𝑓 𝑉𝑉𝑓𝑓 𝑉𝑉𝑓𝑓
𝛾𝛾−1 𝛾𝛾−1
𝑇𝑇𝑖𝑖 𝑉𝑉𝑖𝑖 = 𝑇𝑇𝑓𝑓 𝑉𝑉𝑓𝑓
Adiabatic Processes for an Ideal Gas
• Example-1: A Diesel Engine Cylinder : Air at
20.0°C in the cylinder of a diesel engine is
compressed from an initial pressure of 1.00
atm and volume of 800.0 cm3 to a volume of
60.0 cm3 . (Assume that air behaves as an
ideal gas with 𝛾𝛾 = 1.40 and that the
compression is adiabatic.)
• (a) Find the final pressure and (b) temperature
of the air.
Adiabatic Processes for an Ideal Gas
• Solution: The process is adiabatic since it
happens suddenly. Also, at all time, the gas
behaves like an ideal gas. Hence we have
both:
• 𝑃𝑃𝑉𝑉 𝛾𝛾 = 𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐. and 𝑃𝑃𝑃𝑃 = 𝑛𝑛𝑛𝑛𝑛𝑛 ≠ 𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐.
𝛾𝛾 1.4
𝑉𝑉𝑖𝑖 800.0𝑐𝑐𝑚𝑚3
• 𝑃𝑃𝑓𝑓 = 𝑃𝑃𝑖𝑖 = 1.00 atm
𝑉𝑉𝑓𝑓 60.0 𝑐𝑐𝑚𝑚3
= 37.6 atm
Adiabatic Processes for an Ideal Gas
• Solution (contd.): Because no gas escapes from
the cylinder (𝑛𝑛 = 𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐.) and ideal gas law is
valid, we get
𝑃𝑃𝑖𝑖 𝑉𝑉𝑖𝑖 𝑃𝑃𝑓𝑓 𝑉𝑉𝑓𝑓
• = =>
𝑇𝑇𝑖𝑖 𝑇𝑇𝑓𝑓
𝑃𝑃𝑓𝑓 𝑉𝑉𝑓𝑓 37.6 𝑎𝑎𝑎𝑎𝑎𝑎 60.0𝑐𝑐𝑚𝑚3
𝑇𝑇𝑓𝑓 = 𝑇𝑇𝑖𝑖 = (293 𝐾𝐾)
𝑃𝑃𝑖𝑖 𝑉𝑉𝑖𝑖 1.0 𝑎𝑎𝑎𝑎𝑎𝑎 800.0 𝑐𝑐𝑚𝑚3
𝑜𝑜
= 826𝐾𝐾 = 553 𝐶𝐶
• The temperature is high enough for ignition of
fuel and combustion without the need of spark.
The Equipartition of Energy
• Predictions based on the model for molar specific
heat agree quite well with the behavior of
monatomic gases.
• Molar specific heat (experimental value) does not
agree with theoretical prediction for complex
gases.
• However, for an ideal gas 𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉 = 𝑅𝑅 is the
same for all gases, since the derivation was
independent of the molecular structure.
(based on work done on gas).
The Equipartition of Energy
• For complex molecules,
Internal energy = translational energy KE +
rotational energy KE + vibrational energy
(KE+PE)
The Equipartition of Energy
• For translational motion, we saw:
1
• Each direction (𝑥𝑥, 𝑦𝑦, 𝑧𝑧) contributes
𝑘𝑘𝐵𝐵 𝑇𝑇
2
energy on the average per molecule.
• Theorem of equipartition of energy states
(from statistical mechanics):
𝟏𝟏
Each degree of freedom contributes 𝒌𝒌𝑩𝑩 𝑻𝑻
𝟐𝟐
energy on the average per molecule.
The Equipartition of Energy
• Example-1: Diatomic Molecule:
• Translational motion along 𝑥𝑥, 𝑦𝑦, 𝑧𝑧-directions
• Only two rotational degrees of freedom that
correspond to the rotation about the two
orthogonal axes perpendicular to the bond.
𝑧𝑧̂ 𝑥𝑥� 𝑥𝑥�

𝑦𝑦�
The Equipartition of Energy
• Diatomic Molecule (contd.): We can neglect
the rotation about the y axis because the
molecule’s moment of inertia 𝐼𝐼𝑦𝑦 about its own
axis is small.
1
• Rotational energy 𝐼𝐼𝑦𝑦 𝜔𝜔2
is also small unless
2
at very high temperature (when 𝜔𝜔 is large).
• => Two effective rotational DoFs.
• Energy of rotation about the molecule’s axis is not
accessible at lower temperatures.
The Equipartition of Energy
• Diatomic Molecule (contd.): Two atoms are
joined by an imaginary spring.
• Vibrational motion adds two more degrees of
freedom, which correspond to the kinetic
energy and the potential energy.
• Vibrational energy is accessible
r
only at higher temperatures.
N N

PE = ½K (r – ro)2
The Equipartition of Energy
• Diatomic Molecule (contd.):

Inaccessible Accessible state


states

There are five degrees of freedom accessible at lower temperatures.


The Equipartition of Energy
• Diatomic Molecule (contd.):
• At lower temperatures, total internal energy
𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = (number of molecules)(Average energy
of each molecule)
1 5 5
• 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑁𝑁 3 + 2 𝑘𝑘𝐵𝐵 𝑇𝑇 = 𝑁𝑁𝑘𝑘𝐵𝐵 𝑇𝑇 = 𝑛𝑛𝑛𝑛𝑛𝑛
2 2 2
1 𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 5 7
• 𝐶𝐶𝑉𝑉 = = 𝑅𝑅, 𝐶𝐶𝑃𝑃 = 𝐶𝐶𝑉𝑉 + 𝑅𝑅 = 𝑅𝑅
𝑛𝑛 𝑑𝑑𝑑𝑑 2 2
𝐶𝐶𝑃𝑃 7
• 𝛾𝛾𝑑𝑑𝑑𝑑𝑑𝑑 = = = 1.4 (lower temperatures)
𝐶𝐶𝑉𝑉 5
The Equipartition of Energy
• Diatomic Molecule (contd.):
• At higher temperatures: Two vibrational
degrees of freedom are used;
3+2+2 7 7
• 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑁𝑁 𝑘𝑘𝐵𝐵 𝑇𝑇 = 𝑁𝑁𝑘𝑘𝐵𝐵 𝑇𝑇 = 𝑛𝑛𝑛𝑛𝑛𝑛
2 2 2
1 𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 7 9
• 𝐶𝐶𝑉𝑉 = = 𝑅𝑅, 𝐶𝐶𝑃𝑃 = 𝐶𝐶𝑉𝑉 + 𝑅𝑅 = 𝑅𝑅
𝑛𝑛 𝑑𝑑𝑑𝑑 2 2
𝐶𝐶𝑃𝑃 9
• 𝛾𝛾𝑑𝑑𝑑𝑑𝑑𝑑 = = = 1.286 (higher temperatures)
𝐶𝐶𝑉𝑉 7
The Equipartition of Energy
• Diatomic Molecule (contd.): If measurements
of molar specific heat are made over a wide
temperature range, we see change of 𝐶𝐶𝑉𝑉 , 𝐶𝐶𝑃𝑃
and 𝛾𝛾.

Molar Specific heat of 𝐻𝐻2


The Equipartition of Energy
• At low temperatures, the energy that a molecule
gains in collisions with its neighbors is generally
not large enough to raise it to the first excited
state of either rotation or vibration.
• Thus, even though rotation and vibration are
classically allowed, they do not occur at low
temperatures.
• All molecules are in the ground state for rotation
and vibration at low temperatures. No
contribution at room temperature from vibration.
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
The Equipartition of Energy
• Diatomic Molecule (contd.): If measurements
of molar specific heat are made over a wide
temperature range, we see change of 𝐶𝐶𝑉𝑉 , 𝐶𝐶𝑃𝑃
and 𝛾𝛾.

Molar Specific heat of 𝐻𝐻2


The Equipartition of Energy
• At low temperatures, the energy that a molecule
gains in collisions with its neighbors is generally
not large enough to raise it to the first excited
state of either rotation or vibration.
• Thus, even though rotation and vibration are
classically allowed, they do not occur at low
temperatures.
• All molecules are in the ground state for rotation
and vibration at low temperatures. No
contribution at room temperature from vibration.
The Equipartition of Energy
• Quiz-1: The molar specific heat of a diatomic gas
is measured at constant volume and found to be
29.1 J/mol.K (𝑅𝑅 = 8.31 J/mol.K). The types of
energy that are contributing to the molar specific
heat are
• (a) translation only
• (b) translation and rotation only
• (c) translation and vibration only
• (d) translation, rotation, and vibration.
7
• Ans: 𝐶𝐶𝑉𝑉 = 𝑅𝑅, (d)
2
The Equipartition of Energy
• Quiz-2: The molar specific heat of a gas is
measured at constant volume and found to be
11
𝑅𝑅. The gas is most likely to be
2
• (a) monatomic
• (b) diatomic
• (c) polyatomic.
• Ans: ployatomic
Quiz
• 1. A 1.00-mol sample of a diatomic ideal gas has pressure
𝑃𝑃 and volume 𝑉𝑉. When the gas is heated, its pressure triples
and its volume doubles. This heating process includes two
steps, the first at constant pressure and the second at
constant volume. Determine the amount of energy
transferred to the gas by heat. (Diatomic ideal gas has
𝐶𝐶𝑉𝑉 = 5𝑅𝑅/2 and 𝐶𝐶𝑃𝑃 = 7𝑅𝑅/2.)

• Solution:
• 𝑄𝑄 = 𝑄𝑄𝑉𝑉 + 𝑄𝑄𝑃𝑃 = 𝑛𝑛𝐶𝐶𝑉𝑉 Δ𝑇𝑇 + 𝑛𝑛 𝐶𝐶𝑃𝑃 Δ𝑇𝑇
Quiz- Solution
• For constant pressure, 𝑃𝑃 is constant but
𝑉𝑉𝑖𝑖 → 𝑉𝑉𝑓𝑓𝑎𝑎 = 2 𝑉𝑉𝑖𝑖 , 𝑇𝑇 → 𝑇𝑇𝑓𝑓𝑓𝑓 = 𝑇𝑇𝑖𝑖 + Δ𝑇𝑇 𝑃𝑃
• After this process 𝑎𝑎 : 𝑃𝑃𝑓𝑓𝑓𝑓 = 𝑃𝑃𝑖𝑖 , 𝑉𝑉𝑓𝑓𝑓𝑓 = 2𝑉𝑉𝑖𝑖 , 𝑇𝑇 =
𝑇𝑇𝑓𝑓𝑓𝑓
1 1 1
• Δ𝑇𝑇 𝑃𝑃 = 𝑃𝑃𝑖𝑖 Δ𝑉𝑉 = 𝑃𝑃𝑖𝑖 (𝑉𝑉𝑓𝑓𝑓𝑓 −𝑉𝑉𝑖𝑖 ) = 𝑃𝑃𝑖𝑖 𝑉𝑉𝑖𝑖 =
𝑛𝑛𝑛𝑛 𝑛𝑛𝑛𝑛 𝑛𝑛𝑛𝑛
𝑇𝑇𝑖𝑖 ,
• 𝑇𝑇𝑓𝑓𝑓𝑓 = 𝑇𝑇𝑖𝑖 + Δ𝑇𝑇 𝑃𝑃 = 2𝑇𝑇𝑖𝑖
7 7
• 𝑄𝑄𝑃𝑃 = 𝑛𝑛𝐶𝐶𝑃𝑃 Δ𝑇𝑇 𝑃𝑃 = 𝑛𝑛 𝑅𝑅 𝑇𝑇𝑖𝑖 = 𝑃𝑃𝑃𝑃
2 2
Quiz- Solution
• For constant volume process 𝑏𝑏 , 𝑉𝑉 is constant but
𝑃𝑃𝑖𝑖 → 𝑃𝑃𝑓𝑓𝑓𝑓 = 3𝑃𝑃𝑖𝑖 , 𝑉𝑉 → 𝑉𝑉𝑓𝑓𝑓𝑓 = 𝑉𝑉𝑖𝑖𝑖𝑖 = 𝑉𝑉𝑓𝑓𝑓𝑓 , 𝑇𝑇𝑖𝑖𝑖𝑖 = 𝑇𝑇𝑓𝑓𝑓𝑓 → 𝑇𝑇𝑓𝑓𝑓𝑓
• After this process 𝑏𝑏: 𝑃𝑃𝑓𝑓 = 𝑃𝑃𝑓𝑓𝑓𝑓 = 3𝑃𝑃𝑖𝑖 , 𝑉𝑉 = 𝑉𝑉𝑓𝑓 = 𝑉𝑉𝑓𝑓𝑓𝑓 = 2𝑉𝑉𝑖𝑖
1 1 1
• Δ𝑇𝑇 𝑉𝑉 = 𝑉𝑉Δ𝑃𝑃 = 𝑉𝑉
𝑃𝑃𝑓𝑓 − 𝑃𝑃𝑖𝑖 = (2𝑉𝑉𝑖𝑖 )(2𝑃𝑃𝑖𝑖 ) =
𝑛𝑛𝑛𝑛 𝑖𝑖𝑏𝑏 𝑛𝑛𝑛𝑛 𝑓𝑓𝑓𝑓 𝑛𝑛𝑛𝑛
4𝑇𝑇𝑖𝑖
• 𝑇𝑇𝑓𝑓 = 𝑇𝑇𝑖𝑖𝑖𝑖 + Δ𝑇𝑇 𝑉𝑉 = 2𝑇𝑇𝑖𝑖 + 4𝑇𝑇𝑖𝑖 = 6𝑇𝑇𝑖𝑖
5 20
• 𝑄𝑄𝑉𝑉 = 𝑛𝑛𝐶𝐶𝑉𝑉 Δ𝑇𝑇 𝑉𝑉 = 𝑛𝑛 𝑅𝑅 4𝑇𝑇𝑖𝑖 = 𝑃𝑃𝑃𝑃
2 2
27
• 𝑄𝑄 = 𝑄𝑄𝑃𝑃 + 𝑄𝑄𝑉𝑉 = 𝑃𝑃𝑃𝑃
2
Kinetic Theory of Gases and Mean
Free path
• If we put an odorous gas at one corner of a
room, it will diffuse to the other parts, but it
usually takes a few seconds to do so.
• However, the RMS speed of gas molecules is
quite large (~100 m/s to ~1000 m/s)
• Thus the molecules do not come straight to
us. They collide with each other and move in a
random zigzag manner.
Kinetic Theory of Gases and Mean
Free path
• Between collisions, molecules move with
constant speed (Newton’s first law) along
straight lines.
• Free Path: The length of the
path between two collisions
of the molecules of a gas is
called a free path. It is the
length of a straight line segment.
Kinetic Theory of Gases and Mean
Free path
• Mean Free Path: The average of free paths is
called the mean free path.
• It is the average distance along
which a molecule of an ideal gas
would travel without colliding
with another molecule.
Kinetic Theory of Gases and Mean
Free path
• The Mean free path should be a function of
(a) How large a molecule is i.e. its diameter
(b) How many molecules are present, i.e.
density of the gas.
• Assume that molecules are spherical in shape
with diameter 𝑑𝑑.
• No two molecules would collide unless their
centers come at or to within a distance of 𝑑𝑑.
Kinetic Theory of Gases and Mean
Free path
• This is equivalent to considering one of the
molecules to be a point-particle and the other
one to have a diameter of 2𝑑𝑑 .

• If the paths of the molecules come within 𝑑𝑑


perpendicular distance, they collide.
Kinetic Theory of Gases and Mean
Free path
• Let us choose the large molecule to be the
moving one, with average speed 𝑣𝑣̅ .
• In time interval Δt, the molecule travels a
distance 𝑣𝑣̅ Δt.
• In this time, the molecule sweeps out a
cylindrical volume with cross-sectional area
𝜋𝜋𝑑𝑑 2 and length 𝑣𝑣̅ Δ𝑡𝑡.
Kinetic Theory of Gases and Mean
Free path
• Volume swept out by the molecule: (𝜋𝜋𝑑𝑑 2 ) 𝑣𝑣̅ Δ𝑡𝑡
• 𝑛𝑛𝑉𝑉 =number density of
molecules = number/volume
• The number of point-sized
molecules in the cylinder, with which the large
molecule collides is 𝜋𝜋𝑑𝑑 2 𝑣𝑣̅ Δ𝑡𝑡 𝑛𝑛𝑉𝑉
• Number of collisions in time interval Δ𝑡𝑡 =
number of molecules in the cylinder = 𝜋𝜋𝑑𝑑 2 𝑣𝑣̅ Δ𝑡𝑡𝑛𝑛𝑉𝑉
Kinetic Theory of Gases and Mean
Free path
• Mean free path of the large molecule= (distance
travelled in time interval)/(number of collisions in
time interval)
𝑣𝑣�Δ𝑡𝑡 1
• 𝑙𝑙 = =
𝜋𝜋𝜋𝜋 2 𝑣𝑣�Δ𝑡𝑡𝑛𝑛𝑉𝑉 𝜋𝜋𝜋𝜋 2 𝑛𝑛𝑉𝑉
• Collision frequency 𝑓𝑓 = number of collisions in
𝜋𝜋𝜋𝜋 2 𝑣𝑣�Δ𝑡𝑡𝑛𝑛𝑉𝑉
unit time = = 𝜋𝜋𝜋𝜋 2 𝑣𝑣̅ 𝑛𝑛𝑉𝑉
Δ𝑡𝑡
• Mean Free Time: Average time between two
1 1
collisions = = 2 �𝑛𝑛
𝑓𝑓 𝜋𝜋𝜋𝜋 𝑣𝑣 𝑉𝑉
Kinetic Theory of Gases and Mean
Free path
• Our analysis has assumed that molecules in
the cylinder are stationary.
• When the motion of these molecules is
included in the calculation, the correct results
are:
1 2 𝑣𝑣�
• 𝑙𝑙 = , 𝑓𝑓 = 2 𝜋𝜋𝜋𝜋 𝑣𝑣̅ 𝑛𝑛𝑉𝑉 =
2 𝜋𝜋𝜋𝜋2 𝑛𝑛𝑉𝑉 𝑙𝑙
Kinetic Theory of Gases and Mean
Free path
• Example: Bouncing Around in the Air
• Approximate the air around you as a collection
of nitrogen molecules, each having a diameter
of 2.00 × 10−10 m.
• (A) How far does a typical molecule move
before it collides with another molecule?
• (B) On average, how frequently does one
molecule collide with another?(𝑣𝑣𝑟𝑟𝑟𝑟𝑟𝑟 =
511m/s)
Kinetic Theory of Gases and Mean
Free path
𝑁𝑁 𝑃𝑃
• Solution: The gas is ideal. 𝑛𝑛𝑉𝑉 = = =
𝑉𝑉 𝑘𝑘𝐵𝐵 𝑇𝑇
1.0𝑎𝑎𝑎𝑎𝑎𝑎 1.013×105 𝑃𝑃𝑃𝑃
−23 =
(1.38×10 𝐽𝐽/𝐾𝐾)(293𝐾𝐾) (1.38×10−23 𝐽𝐽/𝐾𝐾)(293𝐾𝐾)
• 𝑛𝑛𝑉𝑉 = 2.50 × 1025 /𝑚𝑚3
1 1
• 𝑙𝑙 = 2 = −10 = 2.25 ×
2 𝜋𝜋𝜋𝜋 𝑛𝑛𝑉𝑉 2 𝜋𝜋(2.0× 10 𝑚𝑚)𝑛𝑛𝑉𝑉
−7
10 𝑚𝑚
Kinetic Theory of Gases and Mean
Free path
𝑚𝑚 3𝑘𝑘𝐵𝐵 𝑇𝑇 8𝑘𝑘𝐵𝐵 𝑇𝑇 𝑚𝑚
• 𝑣𝑣𝑟𝑟𝑟𝑟𝑟𝑟 = 511 = ⇒ 𝑣𝑣̅ = = 473
𝑠𝑠 𝑚𝑚 𝜋𝜋𝜋𝜋 𝑠𝑠
𝑣𝑣� 473𝑚𝑚/𝑠𝑠
• 𝑓𝑓 = = = 2.10 × 109 /𝑠𝑠
𝑙𝑙 2.25×10−7 𝑚𝑚
• The molecule collides with other molecules at the
average rate of about two billion times each
second!
• Note: The average separation 𝑠𝑠̅ between particles
1

is approximately 𝑛𝑛𝑉𝑉 3
, not 𝑙𝑙. Explain.
Kinetic Theory of Gases and Molar
Specific Heats
1 3
• 𝑚𝑚𝑣𝑣 2 = 𝑘𝑘𝐵𝐵 𝑇𝑇, translational motion
2 2
3
• 𝐾𝐾𝑡𝑡𝑡𝑡𝑡𝑡,𝑡𝑡𝑡𝑡𝑡𝑡 = 𝑛𝑛𝑛𝑛𝑛𝑛
2
• 𝑣𝑣𝑟𝑟𝑟𝑟𝑟𝑟 = 3 𝑘𝑘𝐵𝐵 𝑇𝑇/𝑚𝑚 = 3𝑅𝑅𝑅𝑅/𝑀𝑀
1 𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖
• 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑛𝑛𝐶𝐶𝑉𝑉 𝑇𝑇, 𝐶𝐶𝑉𝑉 =
𝑛𝑛 𝑑𝑑𝑑𝑑
𝐷𝐷
• 𝐶𝐶𝑉𝑉 = 𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉 = 𝑅𝑅 , for ideal gas, 𝐷𝐷 =no. of
𝑅𝑅,
2
degrees of freedom, 𝛾𝛾 = 𝐶𝐶𝑃𝑃 /𝐶𝐶𝑉𝑉 = 𝐷𝐷 + 2 /𝐷𝐷
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Kinetic Theory of Gases and Molar
Specific Heats
1 3
• 𝑚𝑚𝑣𝑣 2 = 𝑘𝑘𝐵𝐵 𝑇𝑇, translational motion
2 2
3
• 𝐾𝐾𝑡𝑡𝑡𝑡𝑡𝑡,𝑡𝑡𝑡𝑡𝑡𝑡 = 𝑛𝑛𝑛𝑛𝑛𝑛
2
• 𝑣𝑣𝑟𝑟𝑟𝑟𝑟𝑟 = 3 𝑘𝑘𝐵𝐵 𝑇𝑇/𝑚𝑚 = 3𝑅𝑅𝑅𝑅/𝑀𝑀
1 𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖
• 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑛𝑛𝐶𝐶𝑉𝑉 𝑇𝑇, 𝐶𝐶𝑉𝑉 =
𝑛𝑛 𝑑𝑑𝑑𝑑
𝐷𝐷
• 𝐶𝐶𝑉𝑉 = 𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉 = 𝑅𝑅 , for ideal gas, 𝐷𝐷 =no. of
𝑅𝑅,
2
degrees of freedom, 𝛾𝛾 = 𝐶𝐶𝑃𝑃 /𝐶𝐶𝑉𝑉 = 𝐷𝐷 + 2 /𝐷𝐷
Boltzmann Distribution Law
• We considered only average values of the
energies of molecules in a gas.
• Not all molecules have the same speed or energy.
There is a distribution of speeds and hence of the
energies among the molecules.
• The motion of the molecules is extremely chaotic.
• Any individual molecule is colliding with others at
an enormous rate—typically, a billion times per
second.
Boltzmann Distribution Law
• Each collision results in a change in the speed
and direction of motion.
1 1 2 3
• We have, 𝐾𝐾. 𝐸𝐸. = 𝑚𝑚𝑣𝑣 2 = 𝑚𝑚 𝑣𝑣𝑟𝑟𝑟𝑟𝑟𝑟 = 𝑘𝑘𝐵𝐵 𝑇𝑇,
2 2 2
• Hence, RMS molecular speeds increase with
increasing temperature.
• Define, 𝑛𝑛𝐸𝐸 𝑑𝑑𝑑𝑑 = 𝑛𝑛𝐸𝐸 𝐸𝐸 𝑑𝑑𝑑𝑑 =number of
molecules per unit volume, in the energy
interval between 𝐸𝐸 to 𝐸𝐸 + 𝑑𝑑𝑑𝑑.
Boltzmann Distribution Law
• 𝑛𝑛𝐸𝐸 (𝐸𝐸) = number of molecules per unit
volume, per unit energy interval, interval
between 𝐸𝐸 to 𝐸𝐸 + 𝑑𝑑𝑑𝑑.
• From statistical mechanics:
• 𝑛𝑛𝐸𝐸 𝐸𝐸 = 𝑛𝑛0 𝑒𝑒 −𝐸𝐸/𝑘𝑘𝐵𝐵 𝑇𝑇 , where 𝑒𝑒 −𝐸𝐸/𝑘𝑘𝐵𝐵 𝑇𝑇 is called
the Boltzmann factor.
• 𝑛𝑛0 𝑑𝑑𝑑𝑑 = number of molecules per unit volume
having energies between 𝐸𝐸 = 0 and
𝐸𝐸 = 0 + 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 . i.e. 𝑛𝑛0 = 𝑛𝑛𝑉𝑉 (𝐸𝐸 = 0)
Boltzmann Distribution Law
• Boltzmann Distribution Law: The probability
of finding the molecules in a particular
energy state varies exponentially as the
negative of the energy divided by 𝒌𝒌𝑩𝑩 𝑻𝑻.
• Probability: The ratio of the number of
molecules that have the desired characteristic
to the total number of molecules is the
probability that a particular molecule has that
characteristic.
Boltzmann Distribution Law
• We define:
• 𝑛𝑛𝐸𝐸 𝐸𝐸 𝑑𝑑𝑑𝑑/𝑁𝑁 =probability of having energy in
between 𝐸𝐸 to 𝐸𝐸 + 𝑑𝑑𝑑𝑑.
• 𝑁𝑁𝑣𝑣 𝑣𝑣 𝑑𝑑𝑑𝑑/𝑁𝑁 = probability of having a molecule
speed between 𝑣𝑣 and 𝑣𝑣 + 𝑑𝑑𝑑𝑑
• But 𝑛𝑛𝐸𝐸 𝐸𝐸 𝑑𝑑𝑑𝑑/𝑁𝑁 = 𝑁𝑁𝑣𝑣 𝑣𝑣 𝑑𝑑𝑑𝑑/𝑁𝑁 since, a molecule
1
with speed 𝑣𝑣 has an energy 𝐸𝐸 = 𝑚𝑚𝑣𝑣 2
2
• According to Boltzmann: prob ∝ 𝑒𝑒 −𝐸𝐸/𝑘𝑘𝐵𝐵 𝑇𝑇 .
Maxwell-Boltzmann Distribution
• James Clerk Maxwell (1831–1879) derived an
expression in 1860 that describes the
distribution of molecular speeds.
• Intuitively, we expect that the speed
distribution depends on temperature.
• We expect that the distribution peaks in the
vicinity of 𝑣𝑣𝑟𝑟𝑟𝑟𝑟𝑟 at any temperature, because
𝑣𝑣𝑟𝑟𝑟𝑟𝑟𝑟 is related with temperature.
Maxwell-Boltzmann Distribution
• Maxwell–Boltzmann speed distribution function,
𝑁𝑁𝑣𝑣 (𝑣𝑣) is defined as (from statistical mechanics):
3
𝑚𝑚 2
2 −𝑚𝑚𝑣𝑣 2 /2𝑘𝑘𝐵𝐵 𝑇𝑇
𝑁𝑁𝑣𝑣 𝑣𝑣 = 4𝜋𝜋𝜋𝜋 𝑣𝑣 𝑒𝑒
2 𝜋𝜋𝑘𝑘𝐵𝐵 𝑇𝑇
• where , 𝑁𝑁𝑣𝑣 𝑣𝑣 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 = the number of
molecules with speeds between 𝑣𝑣 and 𝑣𝑣 + 𝑑𝑑𝑑𝑑
• Here, 𝑁𝑁 = total number of molecules in an
ensemble.
Maxwell-Boltzmann Distribution
• Thus, the fraction of molecules with speeds
between 𝑣𝑣 and 𝑣𝑣 + 𝑑𝑑𝑑𝑑 is = 𝑁𝑁𝑣𝑣 𝑣𝑣 𝑑𝑑𝑑𝑑/𝑁𝑁
• 𝑁𝑁𝑣𝑣 𝑣𝑣 𝑑𝑑𝑑𝑑/𝑁𝑁 =Probability of a molecule to
have speed between 𝑣𝑣 and 𝑣𝑣 + 𝑑𝑑𝑑𝑑
• Thus, Probability of a molecule to have speed
between 𝑣𝑣 and 𝑣𝑣 + 𝑑𝑑𝑑𝑑 , per unit speed
interval = (𝑁𝑁𝑣𝑣 𝑣𝑣 𝑑𝑑𝑑𝑑/𝑁𝑁)/ 𝑑𝑑𝑑𝑑 = 𝑁𝑁𝑣𝑣 𝑣𝑣 /𝑁𝑁= 𝑓𝑓 𝑣𝑣
• 𝑓𝑓 𝑣𝑣 = probability /(volume×speed) =
probability density
Maxwell-Boltzmann Distribution
3
𝑚𝑚 2
2 −𝑚𝑚𝑣𝑣 2 /2𝑘𝑘𝐵𝐵 𝑇𝑇
𝑁𝑁𝑣𝑣 𝑣𝑣 = 4 𝜋𝜋𝜋𝜋 𝑣𝑣 𝑒𝑒
2 𝜋𝜋𝑘𝑘𝐵𝐵 𝑇𝑇
1
• Here, 𝐸𝐸 = 𝑚𝑚𝑣𝑣 2 =K.E. of one molecule with
2
speed 𝑣𝑣.
−𝑚𝑚𝑣𝑣 2 /2𝑘𝑘𝐵𝐵 𝑇𝑇
• 𝑒𝑒 = 𝑒𝑒 −𝐸𝐸/𝑘𝑘𝐵𝐵 𝑇𝑇 =Boltzmann factor
• 𝑓𝑓 𝑣𝑣 = 𝑁𝑁𝑣𝑣 𝑣𝑣 /𝑁𝑁, probability density =
probability per unit interval of speed
Maxwell-Boltzmann Distribution
3
𝑚𝑚 2
2 −𝑚𝑚𝑣𝑣 2 /2𝑘𝑘𝐵𝐵 𝑇𝑇
𝑁𝑁𝑣𝑣 𝑣𝑣 = 4 𝜋𝜋𝜋𝜋 𝑣𝑣 𝑒𝑒
2 𝜋𝜋𝑘𝑘𝐵𝐵 𝑇𝑇
• Probability of a molecule to have speed
between 𝑣𝑣 and 𝑣𝑣 + 𝑑𝑑𝑑𝑑
= 𝑓𝑓 𝑣𝑣 𝑑𝑑𝑑𝑑 = 𝑁𝑁𝑣𝑣 𝑣𝑣 𝑑𝑑𝑑𝑑/𝑁𝑁
∞ ∞
• ∫0 𝑓𝑓 𝑣𝑣 𝑑𝑑𝑑𝑑 = ∫0 𝑁𝑁𝑣𝑣 𝑣𝑣 𝑑𝑑𝑑𝑑/𝑁𝑁

∞ ∫0 𝑁𝑁𝑣𝑣 𝑣𝑣 𝑑𝑑𝑑𝑑
⇒ ∫0 𝑑𝑑𝑑𝑑 = =1
𝑁𝑁
Maxwell-Boltzmann Distribution
• Number of Particles:
• Number of gas molecules between speeds 𝑣𝑣
and 𝑣𝑣 + 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 = 𝑁𝑁𝑣𝑣 𝑣𝑣 𝑑𝑑𝑑𝑑
• Total number of particles:

𝑁𝑁 = ∫0 𝑁𝑁𝑣𝑣 𝑣𝑣 𝑑𝑑𝑑𝑑
Molecular Speeds
• Average Speed:
1 𝑁𝑁 ∞ ∞
𝑣𝑣̅ = ∑𝑖𝑖=1 𝑣𝑣𝑖𝑖 = ∫𝑣𝑣=0 𝑓𝑓 𝑣𝑣 𝑣𝑣 𝑑𝑑𝑑𝑑/ ∫𝑣𝑣=0 𝑓𝑓 𝑣𝑣 𝑑𝑑𝑑𝑑
𝑁𝑁

• 𝑣𝑣 𝑑𝑑𝑑𝑑 = 1 =total probability of having
∫𝑣𝑣=0 𝑓𝑓
any value of speed between 0 to ∞.

• 𝑣𝑣̅ = ∫0 𝑣𝑣 𝑁𝑁𝑣𝑣 𝑣𝑣 𝑑𝑑𝑑𝑑 /𝑁𝑁
3

𝑚𝑚 2
2 −𝑚𝑚𝑣𝑣 2 /2𝑘𝑘𝐵𝐵 𝑇𝑇
= 4𝜋𝜋 � 𝑣𝑣 𝑑𝑑𝑑𝑑 𝑣𝑣 𝑒𝑒
2 𝜋𝜋𝑘𝑘𝐵𝐵 𝑇𝑇 0
Molecular Speeds
• Average Speed:
3 3
𝑚𝑚 2 ∞ 3 −𝛼𝛼𝑣𝑣 2 𝑚𝑚 2 1
• 𝑣𝑣̅ = 4𝜋𝜋 ∫0 𝑣𝑣 𝑒𝑒 𝑑𝑑𝑑𝑑 = 4𝜋𝜋
2 𝜋𝜋𝑘𝑘𝐵𝐵 𝑇𝑇 2 𝜋𝜋𝑘𝑘𝐵𝐵 𝑇𝑇 2𝛼𝛼2
3
2 1/2
2 𝑚𝑚 2 2𝑘𝑘𝐵𝐵 𝑇𝑇 2 2𝑘𝑘𝐵𝐵 𝑇𝑇 8𝑘𝑘𝐵𝐵 𝑇𝑇
= = =
𝜋𝜋 2𝑘𝑘𝐵𝐵 𝑇𝑇 𝑚𝑚 𝜋𝜋 𝑚𝑚 𝜋𝜋𝜋𝜋
∞ 3 −𝛼𝛼𝛼𝛼 2 1
• Here we have used ∫0 𝑣𝑣 𝑒𝑒 𝑑𝑑𝑣𝑣 =
2𝛼𝛼2

8𝑘𝑘𝐵𝐵 𝑇𝑇
𝑣𝑣̅ = 𝑣𝑣𝑎𝑎𝑎𝑎𝑎𝑎 =
𝜋𝜋𝜋𝜋
Gaussian Integrals
∞ −𝑥𝑥 2 ∞ −𝑥𝑥 2 𝜋𝜋
• ∫−∞ 𝑒𝑒 𝑑𝑑𝑑𝑑 = 𝜋𝜋 ⇒ ∫0 𝑒𝑒 𝑑𝑑𝑑𝑑 =
2
∞ −𝑠𝑠2
• Proof: Let 𝐼𝐼 = ∫−∞ 𝑒𝑒 𝑑𝑑𝑑𝑑,
2 ∞ −𝑥𝑥 2 ∞ −𝑦𝑦 2
𝐼𝐼 = ∫−∞ 𝑒𝑒 𝑑𝑑𝑑𝑑 ∫−∞ 𝑒𝑒 𝑑𝑑𝑦𝑦
∞ 𝑟𝑟=∞,𝜃𝜃=2𝜋𝜋

𝐼𝐼2 − 𝑥𝑥 2 +𝑦𝑦 2 −𝑟𝑟 2


= � 𝑒𝑒 𝑑𝑑𝑑𝑑 𝑑𝑑𝑑𝑑 = � 𝑒𝑒 𝑟𝑟𝑟𝑟𝑟𝑟 𝑑𝑑𝑑𝑑
𝑥𝑥,𝑦𝑦=−∞ 𝑟𝑟=0,𝜃𝜃=0
∞ ∞ −𝑡𝑡 ∞
2 𝑒𝑒 𝑑𝑑𝑑𝑑
𝐼𝐼 2 = � 𝑟𝑟 −𝑟𝑟
𝑒𝑒 𝑑𝑑𝑟𝑟𝑟𝑟𝑟 = 2 𝜋𝜋 � = 𝜋𝜋 −𝑒𝑒 −𝑡𝑡
0 0 2 0
∞ −𝑥𝑥 2 𝜋𝜋
𝐼𝐼 2 = 𝜋𝜋 𝑒𝑒 −𝑡𝑡 0∞ = 𝜋𝜋, Hence ∫0 𝑒𝑒 𝑑𝑑𝑑𝑑 =
2
Gaussian Integrals
∞ −𝛼𝛼𝑥𝑥 2 𝜋𝜋 ∞ −𝛼𝛼𝑥𝑥 2 1 𝜋𝜋
• ∫−∞ 𝑒𝑒 𝑑𝑑𝑑𝑑 = , ∫0 𝑒𝑒 𝑑𝑑𝑑𝑑 =
𝛼𝛼 2 𝛼𝛼

∞ −𝛼𝛼𝑥𝑥 2 ∞ −𝑠𝑠 2 𝜋𝜋
• Proof: ∫−∞ 𝑒𝑒 𝑑𝑑𝑑𝑑 = ∫−∞ 𝑒𝑒 𝑑𝑑𝑑𝑑/ 𝛼𝛼 =
𝛼𝛼

by scaling. Hence proved.


∞ 2𝑛𝑛+1 −𝛼𝛼𝑥𝑥 2
• ∫−∞ 𝑥𝑥 𝑒𝑒 𝑑𝑑𝑑𝑑 = 0, odd function (proved)
∞ 2𝑛𝑛+1 −𝛼𝛼𝑥𝑥 2 𝑛𝑛!
• ∫0 𝑥𝑥 𝑒𝑒 𝑑𝑑𝑑𝑑 = 𝑛𝑛+1
2𝛼𝛼
Gamma Function

• Definition: Γ 𝑧𝑧 = ∫0 𝑒𝑒 −𝑡𝑡 𝑡𝑡 𝑧𝑧−1 𝑑𝑑𝑑𝑑, for 𝑧𝑧 > 0
• Fundamental property: Γ 𝑧𝑧 + 1 = 𝑧𝑧Γ(𝑧𝑧)
• Γ 𝑛𝑛 = 𝑛𝑛 − 1 !
1
• Γ = 𝜋𝜋 from the Gaussian integral.
2
• Let 𝑡𝑡 = 𝛼𝛼𝑥𝑥 2 , ⇒ 𝑑𝑑𝑑𝑑 = 𝛼𝛼𝛼𝛼𝛼𝛼𝛼𝛼𝛼
∞ −𝛼𝛼𝑥𝑥 2 2𝑛𝑛+1 ∞ 𝑒𝑒 −𝑡𝑡 𝑑𝑑𝑑𝑑 𝑡𝑡 𝑛𝑛 1
• ∫0 𝑒𝑒 𝑥𝑥 𝑑𝑑𝑥𝑥 = ∫0
2𝛼𝛼 𝛼𝛼
= Γ(𝑛𝑛 + 1) 𝑛𝑛+1
2𝛼𝛼

−𝛼𝛼𝑥𝑥 2 2𝑛𝑛+1 𝑛𝑛!
⇒ � 𝑒𝑒 𝑥𝑥 𝑑𝑑𝑑𝑑 =
0 2𝛼𝛼 𝑛𝑛+1
∞ 2 1 ∞ 2 1
• Example: ∫0 𝑥𝑥 𝑒𝑒 −𝛼𝛼𝑥𝑥 𝑑𝑑𝑑𝑑 = , ∫0 𝑥𝑥 3 𝑒𝑒 −𝛼𝛼𝑥𝑥 𝑑𝑑𝑑𝑑 =
2𝛼𝛼 2𝛼𝛼 2
Gaussian Integrals
∞ 2 −𝛼𝛼𝑥𝑥 2 1 𝜋𝜋
• ∫0 𝑥𝑥 𝑒𝑒 𝑑𝑑𝑑𝑑 =
4𝛼𝛼 𝛼𝛼
∞ −𝛼𝛼𝑥𝑥 2 𝜋𝜋
• Proof: We have ∫−∞ 𝑒𝑒 𝑑𝑑𝑑𝑑 = Differentiating
𝛼𝛼
w.r.t 𝛼𝛼,
𝜕𝜕 ∞ −𝛼𝛼𝑥𝑥 2 𝜕𝜕 𝜋𝜋 𝜋𝜋
• ∫ 𝑒𝑒 𝑑𝑑𝑑𝑑 = =−
𝜕𝜕𝛼𝛼 −∞ 𝜕𝜕𝜕𝜕 𝛼𝛼 2𝛼𝛼3/2
∞ 2 −𝛼𝛼𝑥𝑥 2 𝜋𝜋
• ∫−∞ (−𝑥𝑥 )𝑒𝑒 𝑑𝑑𝑑𝑑 = −
2𝛼𝛼3/2
∞ 2 −𝛼𝛼𝑥𝑥 2 𝜋𝜋
• ∫0 𝑥𝑥 𝑒𝑒 𝑑𝑑𝑑𝑑 = 3/2 (proved).
4𝛼𝛼
Gaussian Integrals
∞ 4 −𝛼𝛼𝑥𝑥 2 3 𝜋𝜋
• ∫0 𝑥𝑥 𝑒𝑒 𝑑𝑑𝑑𝑑 =
8𝛼𝛼 2 𝛼𝛼
∞ 2 −𝛼𝛼𝑥𝑥 2 𝜋𝜋
• Proof: We have ∫−∞ 𝑥𝑥 𝑒𝑒 𝑑𝑑𝑑𝑑 =
4𝛼𝛼3/2
Differentiating w.r.t 𝛼𝛼
𝜕𝜕 ∞ 2 −𝛼𝛼𝑥𝑥 2 𝜕𝜕 𝜋𝜋 3 𝜋𝜋
• ∫ 𝑥𝑥 𝑒𝑒 𝑑𝑑𝑑𝑑 = = − 5/2
𝜕𝜕𝜕𝜕 −∞ 𝜕𝜕𝜕𝜕 4𝛼𝛼 3/2 8𝛼𝛼
∞ 4 −𝛼𝛼𝑥𝑥 2 𝜋𝜋
• ∫−∞ 𝑥𝑥 𝑒𝑒 𝑑𝑑𝑑𝑑 = (proved).
8𝛼𝛼 5/2
Molecular Speeds
• RMS Speed:
1 𝑁𝑁
• 2
𝑣𝑣𝑟𝑟𝑟𝑟𝑟𝑟 = ∑𝑖𝑖=1 𝑣𝑣𝑖𝑖2 , for discrete speeds
𝑁𝑁
2 ∞ 2 ∞
• 𝑣𝑣𝑟𝑟𝑟𝑟𝑟𝑟 = ∫𝑣𝑣=0 𝑓𝑓 𝑣𝑣 𝑣𝑣 𝑑𝑑𝑑𝑑/ ∫𝑣𝑣=0 𝑓𝑓 𝑣𝑣 𝑑𝑑𝑑𝑑
2
• 𝑣𝑣𝑟𝑟𝑟𝑟𝑟𝑟 =
3 3
𝑚𝑚 2 ∞ 2 𝑚𝑚 2 3 𝜋𝜋
4𝜋𝜋 ∫0 𝑣𝑣 4 𝑒𝑒 −𝛼𝛼𝑣𝑣 𝑑𝑑𝑑𝑑 = 4𝜋𝜋
2 𝜋𝜋𝑘𝑘𝐵𝐵 𝑇𝑇 2 𝜋𝜋𝑘𝑘𝐵𝐵 𝑇𝑇 8𝛼𝛼5/2
3
5/2
3 𝑚𝑚 2 2𝑘𝑘𝐵𝐵 𝑇𝑇 3 2𝑘𝑘𝐵𝐵 𝑇𝑇 3𝑘𝑘𝐵𝐵 𝑇𝑇
= = =
2 2𝑘𝑘𝐵𝐵 𝑇𝑇 𝑚𝑚 2 𝑚𝑚 𝑚𝑚
Molecular Speeds
• RMS speed:
3𝑘𝑘𝐵𝐵 𝑇𝑇
• 𝑣𝑣𝑟𝑟𝑟𝑟𝑟𝑟 =
𝑚𝑚
𝜕𝜕
• Most Probable Speed: 𝑓𝑓 𝑣𝑣 |𝑣𝑣𝑝𝑝 = 0
𝜕𝜕𝑣𝑣
3
𝜕𝜕 𝑚𝑚 2
2
−𝑚𝑚𝑣𝑣 2 /2𝑘𝑘 𝑇𝑇 𝜕𝜕 2 −𝛼𝛼𝑣𝑣 2
4𝜋𝜋 𝑣𝑣 𝑒𝑒 𝐵𝐵 = 𝛽𝛽𝑣𝑣 𝑒𝑒 � =0
𝜕𝜕𝜕𝜕 2 𝜋𝜋𝑘𝑘𝐵𝐵 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑣𝑣𝑝𝑝
−𝛼𝛼𝑣𝑣𝑝𝑝2 2 −𝛼𝛼𝑣𝑣𝑝𝑝2
⇒ 2𝛽𝛽𝑣𝑣𝑝𝑝 𝑒𝑒 − 𝛽𝛽𝑣𝑣𝑝𝑝 𝛼𝛼𝛼𝑣𝑣𝑝𝑝 𝑒𝑒 =0
1 1 𝑘𝑘𝐵𝐵 𝑇𝑇
• 𝑣𝑣𝑝𝑝2 = ⇒ 𝑣𝑣𝑝𝑝 = = 2𝑘𝑘𝐵𝐵 𝑇𝑇/𝑚𝑚 = 1.4
𝛼𝛼 𝛼𝛼 𝑚𝑚
Molecular Speeds
• Relation between speeds: We see that
3𝑘𝑘𝐵𝐵 𝑇𝑇 8𝑘𝑘𝐵𝐵 𝑇𝑇 2𝑘𝑘𝐵𝐵 𝑇𝑇
𝑣𝑣𝑟𝑟𝑟𝑟𝑟𝑟 > 𝑣𝑣̅ > 𝑣𝑣𝑝𝑝 i.e. > >
𝑚𝑚 𝜋𝜋𝜋𝜋 𝑚𝑚
• Example: 𝑁𝑁 = 105 N molecules
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Quiz-2
• Starting from the first law of thermodynamics,
show that for a quasi-static adiabatic process,
the work done on an ideal gas is given by
𝑊𝑊 = (𝑃𝑃𝑓𝑓 𝑉𝑉𝑓𝑓 − 𝑃𝑃𝑖𝑖 𝑉𝑉𝑖𝑖 )/(𝛾𝛾 − 1), where the
symbols have their usual meanings.
Quiz-2
• Solution: From first law: Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄 + 𝑊𝑊 = 𝑄𝑄𝑖𝑖𝑖𝑖 +
𝑊𝑊𝑜𝑜𝑜𝑜 , In the infinitesimal form:
𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑑𝑑𝑄𝑄𝑖𝑖𝑖𝑖 + 𝑑𝑑𝑊𝑊𝑜𝑜𝑜𝑜
• For an ideal gas, 𝑃𝑃𝑃𝑃 = 𝑛𝑛𝑛𝑛𝑛𝑛,
• For adiabatic process: 𝑑𝑑𝑄𝑄𝑖𝑖𝑖𝑖 = 0
• 𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑑𝑑𝑊𝑊𝑜𝑜𝑜𝑜 = −𝑃𝑃𝑃𝑃𝑃𝑃
𝑇𝑇𝑓𝑓
• But again 𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑛𝑛𝑛𝑛𝑉𝑉 𝑑𝑑𝑑𝑑 ⇒ 𝑊𝑊𝑜𝑜𝑜𝑜 = 𝑛𝑛𝐶𝐶𝑉𝑉 ∫𝑇𝑇 𝑑𝑑𝑑𝑑
𝑖𝑖
𝑃𝑃𝑓𝑓 𝑉𝑉𝑓𝑓 𝑃𝑃𝑖𝑖 𝑉𝑉𝑖𝑖
• 𝑊𝑊𝑜𝑜𝑜𝑜 = 𝑛𝑛𝐶𝐶𝑉𝑉 𝑇𝑇𝑓𝑓 − 𝑇𝑇𝑖𝑖 = 𝑛𝑛𝐶𝐶𝑉𝑉 −
𝑛𝑛𝑛𝑛 𝑛𝑛𝑛𝑛
Quiz-2
• Solution (cntd.):
𝐶𝐶𝑉𝑉
• 𝑊𝑊𝑜𝑜𝑜𝑜 = (𝑃𝑃𝑓𝑓 𝑉𝑉𝑓𝑓 − 𝑃𝑃𝑖𝑖 𝑉𝑉𝑖𝑖 ),
𝑅𝑅
• But 𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉 = 𝑅𝑅 ⇒ 𝐶𝐶𝑉𝑉 = 𝐶𝐶𝑃𝑃 − 𝑅𝑅
𝐶𝐶𝑃𝑃 𝑅𝑅 𝑅𝑅 𝑅𝑅
• 1= − = 𝛾𝛾 − ⇒ = 𝛾𝛾 − 1
𝐶𝐶𝑉𝑉 𝐶𝐶𝑉𝑉 𝐶𝐶𝑉𝑉 𝐶𝐶𝑉𝑉
𝑃𝑃𝑓𝑓 𝑉𝑉𝑓𝑓 −𝑃𝑃𝑖𝑖 𝑉𝑉𝑖𝑖
• 𝑊𝑊𝑜𝑜𝑜𝑜 = Proved.
𝛾𝛾−1
Maxwell-Boltzmann Distribution
3
𝑚𝑚 2
2 −𝑚𝑚𝑣𝑣 2 /2𝑘𝑘𝐵𝐵 𝑇𝑇
𝑁𝑁𝑣𝑣 𝑣𝑣 = 4 𝜋𝜋𝜋𝜋 𝑣𝑣 𝑒𝑒
2 𝜋𝜋𝑘𝑘𝐵𝐵 𝑇𝑇
1
• Here, 𝐸𝐸 = 𝑚𝑚𝑣𝑣 2 =K.E. of one molecule with speed 𝑣𝑣.
2
−𝑚𝑚𝑣𝑣 2 /2𝑘𝑘 𝑇𝑇
• 𝑒𝑒 𝐵𝐵 = 𝑒𝑒 −𝐸𝐸/𝑘𝑘𝐵𝐵 𝑇𝑇
=Boltzmann factor
• Probability of a molecule to have speed between 𝑣𝑣 and
𝑣𝑣 + 𝑑𝑑𝑑𝑑 = 𝑓𝑓 𝑣𝑣 𝑑𝑑𝑑𝑑 = 𝑁𝑁𝑣𝑣 𝑣𝑣 𝑑𝑑𝑑𝑑/𝑁𝑁
• 𝑓𝑓 𝑣𝑣 = 𝑁𝑁𝑣𝑣 𝑣𝑣 /𝑁𝑁, probability density = probability
per unit interval of speed
Maxwell-Boltzmann Distribution
3
𝑚𝑚 2
2 −𝑚𝑚𝑣𝑣 2 /2𝑘𝑘𝐵𝐵 𝑇𝑇
𝑁𝑁𝑣𝑣 𝑣𝑣 = 4 𝜋𝜋𝜋𝜋 𝑣𝑣 𝑒𝑒
2 𝜋𝜋𝑘𝑘𝐵𝐵 𝑇𝑇
• 𝑓𝑓 𝑣𝑣 = 𝑁𝑁𝑣𝑣 𝑣𝑣 /𝑁𝑁, probability density =
probability per unit interval of speed
∞ ∞
• ∫0 𝑓𝑓 𝑣𝑣 𝑑𝑑𝑑𝑑 = ∫0 𝑁𝑁𝑣𝑣 𝑣𝑣 𝑑𝑑𝑑𝑑/𝑁𝑁

∞ ∫0 𝑁𝑁𝑣𝑣 𝑣𝑣 𝑑𝑑𝑑𝑑
⇒ ∫0 𝑑𝑑𝑑𝑑 = =1
𝑁𝑁
Maxwell-Boltzmann Distribution
• Number of Particles:
• Number of gas molecules, per unit volume,
having speeds between 𝑣𝑣 and 𝑣𝑣 + 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 =
𝑁𝑁𝑣𝑣 𝑣𝑣 𝑑𝑑𝑑𝑑
• Total number of particles
per unit volume:
∞ ∞
𝑁𝑁 = ∫0 𝑑𝑑𝑑𝑑 = ∫0 𝑁𝑁𝑣𝑣 𝑣𝑣 𝑑𝑑𝑑𝑑
Molecular Speeds
• Average Speed:
1 𝑁𝑁 ∞ ∞
𝑣𝑣̅ = ∑𝑖𝑖=1 𝑣𝑣𝑖𝑖 = ∫𝑣𝑣=0 𝑓𝑓 𝑣𝑣 𝑣𝑣 𝑑𝑑𝑑𝑑/ ∫𝑣𝑣=0 𝑓𝑓 𝑣𝑣 𝑑𝑑𝑑𝑑
𝑁𝑁

• 𝑣𝑣 𝑑𝑑𝑑𝑑 = 1 = total probability of having
∫𝑣𝑣=0 𝑓𝑓
any value of speed between 0 to ∞.

• 𝑣𝑣̅ = ∫0 𝑣𝑣 𝑁𝑁𝑣𝑣 𝑣𝑣 𝑑𝑑𝑑𝑑 /𝑁𝑁
3

𝑚𝑚 2
2 −𝑚𝑚𝑣𝑣 2 /2𝑘𝑘𝐵𝐵 𝑇𝑇
= 4𝜋𝜋 � 𝑣𝑣 𝑑𝑑𝑑𝑑 𝑣𝑣 𝑒𝑒
2 𝜋𝜋𝑘𝑘𝐵𝐵 𝑇𝑇 0
Molecular Speeds
• Average Speed:
3 3
𝑚𝑚 2 ∞ 3 −𝛼𝛼𝑣𝑣 2 𝑚𝑚 2 1
• 𝑣𝑣̅ = 4𝜋𝜋 ∫0 𝑣𝑣 𝑒𝑒 𝑑𝑑𝑑𝑑 = 4𝜋𝜋
2 𝜋𝜋𝑘𝑘𝐵𝐵 𝑇𝑇 2 𝜋𝜋𝑘𝑘𝐵𝐵 𝑇𝑇 2𝛼𝛼2
3
2 1/2
2 𝑚𝑚 2 2𝑘𝑘𝐵𝐵 𝑇𝑇 2 2𝑘𝑘𝐵𝐵 𝑇𝑇 8𝑘𝑘𝐵𝐵 𝑇𝑇
= = =
𝜋𝜋 2𝑘𝑘𝐵𝐵 𝑇𝑇 𝑚𝑚 𝜋𝜋 𝑚𝑚 𝜋𝜋𝜋𝜋
∞ 3 −𝛼𝛼𝛼𝛼 2 1
• Here we have used ∫0 𝑣𝑣 𝑒𝑒 𝑑𝑑𝑣𝑣 =
2𝛼𝛼2

8𝑘𝑘𝐵𝐵 𝑇𝑇
𝑣𝑣̅ = 𝑣𝑣𝑎𝑎𝑎𝑎𝑎𝑎 =
𝜋𝜋𝜋𝜋
Molecular Speeds
• RMS speed:
3𝑘𝑘𝐵𝐵 𝑇𝑇
• 𝑣𝑣𝑟𝑟𝑟𝑟𝑟𝑟 =
𝑚𝑚
𝜕𝜕
• Most Probable Speed: 𝑓𝑓 𝑣𝑣 |𝑣𝑣𝑝𝑝 = 0
𝜕𝜕𝑣𝑣
3
𝜕𝜕 𝑚𝑚 2
2
−𝑚𝑚𝑣𝑣 2 /2𝑘𝑘 𝑇𝑇 𝜕𝜕 2 −𝛼𝛼𝑣𝑣 2
4𝜋𝜋 𝑣𝑣 𝑒𝑒 𝐵𝐵 = 𝛽𝛽𝑣𝑣 𝑒𝑒 � =0
𝜕𝜕𝜕𝜕 2 𝜋𝜋𝑘𝑘𝐵𝐵 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑣𝑣𝑝𝑝
−𝛼𝛼𝑣𝑣𝑝𝑝2 2 −𝛼𝛼𝑣𝑣𝑝𝑝2
⇒ 2𝛽𝛽𝑣𝑣𝑝𝑝 𝑒𝑒 − 𝛽𝛽𝑣𝑣𝑝𝑝 𝛼𝛼𝛼𝑣𝑣𝑝𝑝 𝑒𝑒 =0
1 1 𝑘𝑘𝐵𝐵 𝑇𝑇
• 𝑣𝑣𝑝𝑝2 = ⇒ 𝑣𝑣𝑝𝑝 = = 2𝑘𝑘𝐵𝐵 𝑇𝑇/𝑚𝑚 = 1.4
𝛼𝛼 𝛼𝛼 𝑚𝑚
Molecular Speeds
• Relation between speeds: We see that
3𝑘𝑘𝐵𝐵 𝑇𝑇 8𝑘𝑘𝐵𝐵 𝑇𝑇 2𝑘𝑘𝐵𝐵 𝑇𝑇
𝑣𝑣𝑟𝑟𝑟𝑟𝑟𝑟 > 𝑣𝑣̅ > 𝑣𝑣𝑝𝑝 i.e. > >
𝑚𝑚 𝜋𝜋𝜋𝜋 𝑚𝑚
• Example: 𝑁𝑁 = 105 N molecules
Molecular Speeds
• The speed distribution function for 105 nitrogen
molecules at 300 K and 900 K.
• The total area under either curve is equal to the
total number of molecules, which in this case
equals 105 .
• Quiz-1: Consider the qualitative shapes of the
two curves above. Suppose you have two
containers of gas at the same temperature.
Container A has 105 nitrogen molecules and
container B has 105 hydrogen molecules.
Molecular Speeds
• The correct qualitative matching between the
containers and the two curves is:
• (a) container A corresponds
to the blue curve and container
B to the brown curve
(b) container B corresponds to the blue curve
and container A to the brown curve.
(c) both containers correspond to the same
curve.
Molecular Speeds
• Answer: Because the hydrogen atoms are lighter
than the nitrogen molecules, they move with a
higher average speed and the distribution curve
is stretched out more along the horizontal axis.
• Example-1: Nine particles have speeds of 5.00,
8.00, 12.0, 12.0, 12.0, 14.0, 14.0, 17.0, and 20.0
m/s.
• (a) Find the particles’ average speed.
• (b) What is the rms speed of the particles?
• (c) What is the most probable speed of the
particles?
Molecular Speeds
5+8+12+12+12+14+14+17+20
• Solution: 𝑣𝑣̅ = 𝑚𝑚/𝑠𝑠
9
=12.7 m/s
• 𝑣𝑣𝑟𝑟𝑟𝑟𝑟𝑟 =
52 +82 +122 +122 +122 +14 2 +142 +172 +202
𝑚𝑚/𝑠𝑠
9
• 𝑣𝑣𝑟𝑟𝑟𝑟𝑟𝑟 = 13.3 𝑚𝑚/𝑠𝑠
• 𝑣𝑣𝑝𝑝 = 𝑣𝑣𝑚𝑚𝑚𝑚 = 12.0m/s.
Energy Distribution
• 𝑁𝑁𝐸𝐸 𝐸𝐸 𝑑𝑑𝑑𝑑 = 𝑁𝑁𝑣𝑣 𝑣𝑣 𝑑𝑑𝑑𝑑 ⇒ 𝑁𝑁𝐸𝐸 𝐸𝐸 = 𝑁𝑁𝑣𝑣 𝑣𝑣 𝑑𝑑𝑑𝑑/𝑑𝑑𝑑𝑑
3
𝑚𝑚 2 2 −𝑚𝑚𝑣𝑣 2 /2𝑘𝑘𝐵𝐵 𝑇𝑇
• 𝑁𝑁𝑣𝑣 𝑣𝑣 = 4𝜋𝜋𝜋𝜋 𝑣𝑣 𝑒𝑒
2 𝜋𝜋𝑘𝑘𝐵𝐵 𝑇𝑇
1
2𝑁𝑁 1
• 𝑁𝑁𝐸𝐸 𝐸𝐸 = 3 𝐸𝐸 𝑒𝑒 −𝐸𝐸/𝑘𝑘𝐵𝐵 𝑇𝑇
2
𝜋𝜋
𝑘𝑘𝐵𝐵 𝑇𝑇
2
∞ ∞ 3
• 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝐸𝐸𝑡𝑡𝑡𝑡𝑡𝑡 = ∫0 𝐸𝐸 𝑑𝑑𝑑𝑑 = ∫0 𝐸𝐸𝑁𝑁𝐸𝐸 𝐸𝐸 𝑑𝑑𝑑𝑑 = 𝑁𝑁𝑁𝑁𝐵𝐵 𝑇𝑇
2
3
• �
𝐸𝐸 = 𝐸𝐸𝑡𝑡𝑡𝑡𝑡𝑡 /𝑁𝑁 = 𝑘𝑘𝐵𝐵 𝑇𝑇
2
1 ∞ 3/2 −𝑢𝑢 3
• 𝐸𝐸𝑝𝑝 = 𝑘𝑘𝐵𝐵 𝑇𝑇, prove it. (Use ∫0 𝑢𝑢 𝑒𝑒 𝑑𝑑𝑑𝑑 = 𝜋𝜋)
2 4
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Review of First Law of
Thermodynamics
• The first law of thermodynamics is a
statement of conservation of energy.
• A change in the internal energy in of system
can happen by energy transfer by (a) heat, or
by (b) work done or by (c) both.
• Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄 + 𝑊𝑊,
• 𝑄𝑄 = heat transferred to the system
• 𝑊𝑊 = work done on the system.
Review of First Law of
Thermodynamics
• Both heat transferred and work done depend
on the path taken in the state diagram.
• 𝑄𝑄 + 𝑊𝑊 is independent of the path taken.
• 𝑄𝑄 + 𝑊𝑊 must be related to change of an
intrinsic property of the system.
• 𝑄𝑄 + 𝑊𝑊 must be function of the extreme
states, not of the intermediate states along
the path taken.
Review of First Law of
Thermodynamics
• Makes no distinction between results of heat
transfer and work done.
• State changes may occur by either solely heat
transfer or by work done, solely or by a
combination of both.
• Example: (a) We can move along a constant
volume path and have no work done.
• (b) We can move along an adiabat and have
no heat transfer.
Inadequacies of the First Law
• (a) Makes no distinction between results of
heat transfer and work done.
• (b) Makes no distinction between processes
that occur spontaneously and processes that
do not occur spontaneously (but are still
allowed by the first law).
• (c) All processes that conserve energy are
allowed by the first law.
Inadequacies of the First Law
• There is an important distinction between
heat and work not evident from the first law.
• “It is impossible to design a device that,
operating in a cyclic fashion, takes in energy
by heat and expels an equal amount of energy
by work”.
• But “it is possible to convert work done on a
system in a cyclic process and convert the
whole of it into heat expelled”.
Second Law of Thermodynamics
• Second law states which processes do and
which do not occur.
• Examples: Processes allowed by the first law
but violates second law:
• A. Two objects at different temperatures,
placed in thermal contact, the net transfer of
energy by heat is always from the warmer
object to the cooler object, never from the
cooler to the warmer.
Second Law of Thermodynamics
• B. A rubber ball dropped to the ground
bounces several times and eventually comes
to rest. A ball lying on the ground never
gathers internal energy from the ground and
begins bouncing on its own.
• C. An oscillating pendulum eventually comes
to rest because of collisions with air molecules
and friction at the point of suspension. A
pendulum at rest never starts to oscillate.
Second Law of Thermodynamics
• Reversible Process: A reversible process is one
which can i.e. is allowed to run in the opposite
direction.
• Irreversible Process: An irreversible process
can occur naturally in one direction only.
• No irreversible process has ever been
observed to run backward.
• Restricted by the second law to run in the
opposite direction.
Heat Engines
• Definition: A cyclic device that takes in energy
by heat and expels a fraction of this energy by
work is called a heat engine.
• Ideal Heat Engine: An ideal heat engine is a
cyclic device, that takes in energy by heat and
expels the whole if this energy as work.
• An ideal heat engine converts all the heat
taken in into work done.
Heat Engines
• An ideal heat engine has 100% efficiency, not
possible to implement.
• Examples: Steam engine, internal combustion
engine, etc.
• A heat engine carries a working substance
• The working substance goes through a cyclic
process.
Heat Engines
• During the cycle of a heat engine:
• (a) working substance absorbs energy in the
form of heat from a high-temperature energy
reservoir,
• (b) some fraction of this heat is converted into
work done by the engine
• (c) fraction of the absorbed energy is expelled
in the form of heat to a low temperature
energy reservoir.
Heat Engines
• Example: The Steam Engine: (a) Uses water as
the working substance. Water in a boiler absorbs
energy from burning fuel and evaporates to
steam.
• (b) Steam molecules does work by expanding
against a piston.
• (c) The steam cools and condenses, by expelling
excess heat. The liquid water produced returns to
the boiler and the cycle repeats.
Heat Engines
• The engine absorbs a quantity of energy |𝑄𝑄ℎ |
from the hot reservoir.
• The engine does work 𝑊𝑊𝑒𝑒𝑒𝑒𝑒𝑒 on the
environment ( Work done on the
engine W = −𝑊𝑊𝑒𝑒𝑒𝑒𝑒𝑒 < 0)
• The engine gives up a quantity
of energy |𝑄𝑄𝑐𝑐 | to the cold
reservoir.
Heat Engines
• In a cyclic process Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 0.
• Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄 + 𝑊𝑊 = 0 ⇒ 𝑄𝑄 − 𝑊𝑊𝑒𝑒𝑒𝑒𝑒𝑒 = 0
• 𝑄𝑄 = 𝑊𝑊𝑒𝑒𝑒𝑒𝑒𝑒
The net work done by the heat engine = net
heat transferred to it
• 𝑄𝑄 = 𝑄𝑄𝑛𝑛𝑛𝑛𝑛𝑛 = 𝑄𝑄ℎ − 𝑄𝑄𝑐𝑐
• 𝑊𝑊𝑒𝑒𝑒𝑒𝑒𝑒 = 𝑄𝑄ℎ − 𝑄𝑄𝑐𝑐
Heat Engines
• On a PV diagram: The net work done in a
cyclic process is the area enclosed by the
curve representing the process.
Heat Engines
• Efficiency of a Heat Engine: Thermal efficiency e
of a heat engine
𝑊𝑊𝑒𝑒𝑒𝑒𝑒𝑒 𝑄𝑄ℎ − |𝑄𝑄𝑐𝑐 | |𝑄𝑄𝑐𝑐 |
𝑒𝑒 = = =1−
|𝑄𝑄ℎ | |𝑄𝑄ℎ | |𝑄𝑄ℎ |
• 𝑒𝑒= the ratio of the net work done by the engine
during one cycle to the energy input at the higher
temperature during the cycle
= the ratio of what you gain (work) to what you
give (energy transfer at the higher temperature)
Heat Engines
• All heat engines expel only a fraction of the input
energy |𝑄𝑄ℎ | by mechanical work.
• Thermal efficiency is always less than 100%.
• We have 100% efficiency i.e. (𝑒𝑒 = 1) only if
𝑄𝑄𝑐𝑐 = 0, i.e. no energy is expelled to the cold
reservoir.
• A heat engine with perfect efficiency would have
to convert all of the input energy in the form of
heat to work.
Second Law of Thermodynamics
• In practice, all heat engines have efficieny less
than 1.
It is impossible to construct a heat engine
that, operating in a cycle, produces no effect
other than the input of energy by heat from a
reservoir and the performance of an equal
amount of work.
• Kelvin–Planck form of the second law of
thermodynamics.
Second Law of Thermodynamics
• 𝑊𝑊𝑒𝑒𝑒𝑒𝑒𝑒 can never be equal to |𝑄𝑄ℎ |.
• Some energy |𝑄𝑄𝑄𝑄| must be rejected to the
environment.

• Impossible
engine !
Heat Engines
• Example: The Efficiency of an Engine:
• An engine transfers 2.00 × 103 J of energy
from a hot reservoir during a cycle and
transfers 1.50 × 103 J as exhaust to a cold
reservoir.
• (A) Find the efficiency of the engine.
• (B) How much work does this engine do in one
cycle?
• What is the power output of this engine?
Heat Engines
𝑄𝑄𝑐𝑐 1.50×103 𝐽𝐽
• Solution: e = 1 − =1− = 25%
𝑄𝑄ℎ 2.00×103 𝐽𝐽
• 𝑊𝑊𝑒𝑒𝑒𝑒𝑒𝑒 = 𝑄𝑄ℎ − 𝑄𝑄𝑐𝑐 = 2.00 − 1.50 × 103 𝐽𝐽 =
5.0 × 102 𝐽𝐽
• We can not say anything about the power output.
The power of an engine is the rate at which work
is done by the engine. You know how much work
is done per cycle but you have no information
about the time interval associated with one cycle.
Heat Engines
• Given that the engine operates at 2 000 rpm
(revolutions per minute), we can find the
power output:
𝑊𝑊𝑒𝑒𝑒𝑒𝑒𝑒 5.0×102 𝐽𝐽
• 𝑃𝑃 = = 1 = 1.7 × 104 Watts
𝑇𝑇 ( 𝑚𝑚𝑚𝑚𝑚𝑚)
2000
Heat Pumps or Refrigerators
• In a heat engine, the direction of energy
transfer is from the hot reservoir to the cold
reservoir (thermal direction)
• It processes the energy from the hot reservoir
so as to do useful work.
• How can we transfer energy from the cold
reservoir to the hot reservoir?
• Not the natural direction of energy transfer.
Heat Pumps or Refrigerators
• We must put some energy into a device to
take heat from cold to hot reservoir.
• A device that takes heat from a cold reservoir
to a hot reservoir is called a heat pump or a
refrigerator.
• Example: We cool homes in summer using
heat pumps called air conditioners.
Heat Pumps or Refrigerators
• The engine takes in energy |𝑄𝑄𝑐𝑐 | from a cold
reservoir and expels energy |𝑄𝑄ℎ | to a hot
reservoir.
• This can be accomplished only if work is done on
the engine.
• From first law: energy given up to the hot
reservoir must equal the sum of the work done
and the energy taken in from the cold reservoir.
• 𝑄𝑄ℎ = 𝑄𝑄𝑐𝑐 + 𝑊𝑊𝑜𝑜𝑜𝑜
Heat Pumps or Refrigerators
• The refrigerator or heat pump transfers energy
from a colder body, to a hotter body.
• We would like to carry out this process with a
minimum of work.
• If it could be accomplished without doing any
work, then the refrigerator or heat pump
would be “perfect”
Heat Pumps or Refrigerators
• A perfect refrigerator would be the violation
of the second law of thermodynamics.
It is impossible to construct a cyclical machine
whose sole effect is to transfer energy
continuously by heat from one object to
another object at a higher temperature
without the input of energy by work.
• Clausius statement of the second law of
thermodynamics.
Heat Pumps or Refrigerators
• Real Heat Pump Impossible Heat Pump
Heat Pumps or Refrigerators
• Thus energy does not transfer spontaneously
by heat from a cold object to a hot object.
• This direction of energy transfer requires an
input of energy to a heat pump, which is often
supplied by means of electricity.
• The Clausius and Kelvin–Planck statements of
the second law of thermodynamics appear, at
first sight, to be unrelated, but in fact they are
equivalent in all respects.
Heat Pumps
• The effectiveness of a heat pump is called the
Coefficient of performance (COP)=
energy transferred at high temperature
=
work done by heat pump
|𝑄𝑄ℎ |
=
𝑊𝑊
• Ratio of what you gain (energy delivered to
the interior of the building) to what you give
(work input).
Coefficient of Performance of a Heat
Pump
• COP is similar to the thermal efficiency for a
heat engine.
• Because |𝑄𝑄ℎ | is generally greater than 𝑊𝑊,
typical values for the COP are greater than
unity.
• It is desirable for the COP to be as high as
possible.
• Heat pumps can be used to warm room by
cooling the outside (in colder countries).
Coefficient of Performance of a Heat
Pump
• In the “heating Mode”:
For outside temperature equal to 25∘ F (−4∘ C)
or higher, COP ∼ 4.
• The amount of energy transferred to the building
= 4 × (the work done by the motor in the heat
pump).
• As the outside temperature decreases, it
becomes more difficult for the heat pump to
extract sufficient energy from the air.
Coefficient of Performance of a Heat
Pump
• Hence the COP decreases as outside 𝑇𝑇 drops.
• COP can fall below unity for temperatures
below about 15∘ F (−9∘ C).
• Use heat pumps in colder areas by burying the
external coils deep in the ground.
• Energy is extracted from the ground, which
tends to be warmer than the air outside
during the winter.
Coefficient of Performance for a
Refrigerator
• In “cooling” mode: “what you gain” =
energy removed from the cold reservoir.
• Most efficient refrigerator is the one that
removes the greatest amount of energy from
the cold reservoir in exchange for the least
amount of work.
𝑄𝑄𝑐𝑐
• COP for refrigerator: COP (cooling)=
𝑊𝑊
• Good refrigerator : COP ∼ 5 − 6
Coefficient of Performance
• Example: Freezing Water: A certain refrigerator
has a COP of 5.00. When the refrigerator is
running, its power input is 500 W. A sample of
water of mass 500 g and temperature 20.0°C is
placed in the freezer compartment.
• (A) How long does it take to freeze the water to
ice at 0°C ?
Assume that all other parts of the refrigerator stay at
the same temperature and there is no leakage of energy from
the exterior.
Coefficient of Performance
• Solution: Energy leaves the water, reducing its
temperature and then freezing it into ice.
• Amount of energy that we must extract from 500
g of water at 20°C to turn it into ice at 0°C is:
• |𝑄𝑄𝑐𝑐 | = |𝑚𝑚 𝑐𝑐 Δ𝑇𝑇 + 𝑚𝑚𝐿𝐿𝑓𝑓 |
𝑄𝑄
• 𝑐𝑐 = = specific heat of a substance = heat
𝑚𝑚Δ𝑇𝑇
capacity per unit mass
• 𝐿𝐿𝑓𝑓 = latent heat of freezing
Coefficient of Performance
• Solution (contd.): 𝑐𝑐𝑤𝑤𝑤𝑤𝑤𝑤𝑤𝑤𝑤𝑤 = 4.186J/kg-°C =1.00
cal/g-°C
• 𝐿𝐿𝑓𝑓 = 3.33 × 105 J/kg
• |𝑄𝑄𝑐𝑐 | = |𝑚𝑚 𝑐𝑐 Δ𝑇𝑇 + 𝑚𝑚𝐿𝐿𝑓𝑓 | = 𝑚𝑚 |𝑐𝑐Δ𝑇𝑇 + 𝐿𝐿𝑓𝑓 |
5
4.186𝐽𝐽 ∘
10 𝐽𝐽
= 0.5𝑘𝑘𝑘𝑘 20.0 𝐶𝐶 + 3.33 ×
𝑘𝑘𝑘𝑘 ∘𝐶𝐶 𝑘𝑘𝑘𝑘
= 2.08 × 105 𝐽𝐽
Coefficient of Performance
𝑄𝑄𝑐𝑐 2.08×105 𝐽𝐽
• Solution (contd.): 𝑊𝑊 = = =
𝐶𝐶𝐶𝐶𝐶𝐶 5.00
4.17 × 104 𝐽𝐽= work done in cooling
• Time interval required for the freezing
𝑊𝑊 4.17×104 𝐽𝐽
process: Δ𝑡𝑡 = = = 83.3 𝑠𝑠
𝑃𝑃 500 𝑊𝑊
• This time interval is very different from that of
our everyday experience.
Coefficient of Performance
• Only a small part of the energy extracted from
the refrigerator interior in a given time
interval will come from the water.
• Energy from the interior (present or that leaks
into) must also be extracted.
• In reality, the time interval for the water to
freeze is much longer than 83.3 s.
Reversible and Irreversible Processes
• Reversible Process: In a reversible process,
the system undergoing the process can be
returned to its initial conditions along the
same path on a PV diagram.
• Every point along a reversible path is an
equilibrium state.
• Irreversible Process: In an irreversible process,
a system can not be taken in the backward
direction along its path.
Reversible and Irreversible Processes
• Examples: Irreversible process:
• (a) All natural processes are irreversible
process.
• (b) Adiabatic Free Expansion: A gas in a
thermally insulated container, separated by a
membrane from vacuum.
• When the membrane is punctured, the gas
expands freely into the vacuum.
Reversible and Irreversible Processes
• The system has changed because it occupies a
greater volume and the gas has less pressure.
• No work is done on the gas as the molecules
do not exert force on other molecules;
expands into vacuum.
• No heat transfer occurs.
• The system has changed but
the surroundings have not.
Reversible and Irreversible Processes
• Adiabatic Free Expansion: Irreversible process
• To return the system to original state we need
to:
• (a) Compress the gas (may be by piston-
cylinder mechanism).
• (b) Cool down the gas : Remove the heat
extracted from compression from the system
to cool the gas to its original temperature.
Reversible and Irreversible Processes
• Adiabatic Free Expansion: We could have
used the heat extracted from the compression
to use it to compress the piston.
• Kelvin–Planck statement of the second law
specifies that the energy removed from the
gas cannot be completely converted to
mechanical energy.
• Hence the process is irreversible.
Reversible and Irreversible Processes
• (d) The conduction of heat from a hot object
to a cold one is not reversible.
• (e) Friction can transform work into heat, but
friction can never transform heat into work.
Thus the conversion of work into heat via
friction is not reversible.
• (f) When a system passes through non-
equilibrium states, such as when there is
turbulence in a gas or when a gas explodes.
Reversible and Irreversible Processes
• Examples: Reversible Process:
• Real processes running very slowly, such that
the system is always very nearly in equilibrium
state can be approximated as a reversible
process.
• Gas-cylinder in contact with constant T
reservoir:
• We add small grains of sand to increase the
pressure.
Reversible and Irreversible Processes
• Gas-cylinder in contact with constant T
reservoir:
• As pressure increases, the system
deviates from an equilibrium
state, but is so close to one
that it achieves a new
equilibrium state in a very
short time.
Reversible and Irreversible Processes
• Gas-cylinder in contact with constant T
reservoir: Differences between states are so
small that we can approximate the entire
process as occurring through continuous
equilibrium states.
• The process is quasi-static.
• We can reverse the process
by slowly removing grains from the piston.
Conditions for Reversibility
• Characteristics of a Reversible Process:
• No dissipative effects (such as turbulence or
friction) present.
• Energy transfer as heat can only occur between
objects at the same temperature (or at objects
infinitesimally near the same temperature).
• Process occurs slowly enough i.e. is quasi-static,
for the intermediate steps to be equilibrium
states (or infinitesimally near equilibrium states).
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Carnot Engine
• According to the second law of thermodynamics,
it is impossible for a heat engine working
between two heat reservoirs to be 100%
efficient.
• What, then, is the maximum possible efficiency
for such an engine?
• In 1824, French engineer Sadi Carnot described a
theoretical engine, known as the Carnot engine.
• This was done before either the first or the
second law of thermodynamics had been
established.
Carnot Engine
• Carnot found that a reversible engine is the most
efficient engine that can operate between any
two given reservoirs.
• “A heat engine operating in an ideal, reversible
cycle- called the Carnot cycle- between two
energy reservoirs is the most efficient engine
possible”.
• The Carnot engine is an ideal engine, which
establishes an upper limit on the efficiencies of all
other engines.
Carnot’s Theorem
• Carnot’s theorem states that:
No real heat engine operating between two
energy reservoirs can be more efficient than a
Carnot engine operating between the same
two reservoirs.
• OR, No engine working between two given
heat reservoirs can be more efficient than a
reversible engine (a Carnot engine) working
between those two reservoirs.
Carnot’s Theorem
• Hence, the net work done by a working
substance taken through the Carnot cycle is
the greatest amount of work possible for a
given amount of energy supplied to the
substance at the same higher temperature.
• Proof: Consider two heat engines, operating
between the same two temperatures.
• We will use the method of contradiction to
prove Carnot’s theorem.
Proof of Carnot’s Theorem
• A. Engine-1: Carnot engine with efficiency 𝑒𝑒𝐶𝐶
• B. Engine-2: Another imaginary heat engine
with efficiency 𝑒𝑒 > 𝑒𝑒𝐶𝐶 .
• We run the Carnot engine as a Carnot
refrigerator, to remove heat from the cold
reservoir, at the expense of some work.
• As an illustration, we look at a numerical
example.
Proof of Carnot’s Theorem
• A Carnot engine: A Carnot refrigerator:
Proof of Carnot’s Theorem
• The more efficient engine is used to run the
Carnot refrigerator (it needs some work to run
the refrigerator).
Proof of Carnot’s Theorem
• For the combination of the more efficient
engine and Carnot refrigerator, no exchange
by work with the surroundings occurs.
• The net result of the combination is a transfer
of energy from the cold to the hot reservoir
without work being done on the combination.

+ =
Proof of Carnot’s Theorem
• But, according to the Clausius statement of
the second law, it is impossible to extract heat
from cold reservoir and convert it totally to
work (without any extra energy spent).
Proof of Carnot’s Theorem
• Hence, All real engines are less efficient than
the Carnot engine because they do not
operate through a reversible cycle. (Proved)
• The efficiency of a real engine is further
reduced by such practical difficulties as
friction and energy losses by conduction.
Carnot Cycle
• Consider a Carnot cycle between
temperatures 𝑇𝑇ℎ and 𝑇𝑇𝑐𝑐 .
• The working substance is an ideal gas.
• The gas is contained in a cylinder with a
frictionless movable piston connected to it.
• The cylinder’s walls and the piston are
thermally non-conducting.
Carnot Cycle
• The Carnot cycle consists of two adiabatic
processes and two isothermal processes, all
reversible:
Carnot Cycle
• 1. Process 𝐴𝐴 → 𝐵𝐵: Isothermal expansion at
temperature 𝑇𝑇ℎ .
• 2. Process 𝐵𝐵 → 𝐶𝐶: Adiabatic expansion of the
gas from temperature 𝑇𝑇ℎ to temperature 𝑇𝑇𝑐𝑐 .
• 3. Process 𝐶𝐶 → 𝐷𝐷: Isothermal compression at
temperature 𝑇𝑇𝑐𝑐 .
• 4. Process 𝐷𝐷 → 𝐴𝐴: Adiabatic compression of
the gas from temperature 𝑇𝑇𝑐𝑐 to temperature
𝑇𝑇ℎ .
Carnot Cycle
• Process 𝐴𝐴 → 𝐵𝐵: The gas is in contact with
energy reservoir at 𝑇𝑇ℎ .
• The gas absorbs energy |𝑄𝑄ℎ | from the reservoir
through the base.
• The gas does work 𝑊𝑊𝐴𝐴𝐴𝐴 in raising the piston.
• Process 𝐵𝐵 → 𝐶𝐶: No heat transfer occurs.
• Temperature of the gas decreases from 𝑇𝑇ℎ to 𝑇𝑇𝑐𝑐 .
• The gas does work 𝑊𝑊𝐵𝐵𝐵𝐵 in raising the piston.
Carnot Cycle
• Process 𝐶𝐶 → 𝐷𝐷: The gas in contact with
energy reservoir at 𝑇𝑇𝑐𝑐 .
• The gas expels energy |𝑄𝑄𝑐𝑐 | to the reservoir
through the base. Pressure increases.
• Work done by the piston on the gas is 𝑊𝑊𝐶𝐶𝐶𝐶 .
• Process 𝐷𝐷 → 𝐴𝐴: No heat transfer occurs.
• Temperature of the gas increases to 𝑇𝑇ℎ .
• Work done by the piston on the gas is 𝑊𝑊𝐷𝐷𝐷𝐷 .
Carnot Cycle
Efficiency of a Carnot Cycle
• Net work done in this reversible cyclic process =
area enclosed by the path ABCDA.
• Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 0, ⇒ 𝑄𝑄𝑖𝑖𝑖𝑖 + 𝑊𝑊𝑜𝑜𝑜𝑜 = 0
• 𝑊𝑊𝑒𝑒𝑒𝑒𝑒𝑒 = net work done by the engine
• 𝑄𝑄𝑖𝑖𝑖𝑖 = 𝑄𝑄ℎ − |𝑄𝑄𝑐𝑐 |= net heat energy transferred
into the system.
𝑊𝑊𝑒𝑒𝑒𝑒𝑒𝑒What you gain 𝑄𝑄ℎ −|𝑄𝑄𝑐𝑐 | |Q𝑐𝑐 |
• 𝑒𝑒 = = = =1−
|𝑄𝑄ℎ | What you give |𝑄𝑄ℎ | |𝑄𝑄ℎ |
Efficiency of a Carnot Cycle
• Example: Efficiency of the Carnot Engine
• For isothermal expansion or compressions:
𝑉𝑉𝐵𝐵
• 𝑄𝑄ℎ = −𝑊𝑊𝐴𝐴𝐴𝐴 = 𝑛𝑛𝑛𝑛𝑇𝑇ℎ ln
𝑉𝑉𝐴𝐴
𝑉𝑉𝐶𝐶
• 𝑄𝑄𝑐𝑐 = −𝑊𝑊𝐶𝐶𝐶𝐶 = 𝑛𝑛𝑛𝑛𝑇𝑇𝑐𝑐 ln 0
𝑉𝑉𝐷𝐷
𝑉𝑉
|𝑄𝑄𝑐𝑐 | 𝑇𝑇𝑐𝑐 ln( 𝑐𝑐 )
𝑉𝑉𝐷𝐷
• Hence, dividing we get =
|𝑄𝑄ℎ | 𝑇𝑇ℎ ln(𝑉𝑉𝐵𝐵 /𝑉𝑉𝐴𝐴 )
Efficiency of a Carnot Cycle
• For the two adiabatic processes:
𝛾𝛾−1 𝛾𝛾−1 𝛾𝛾−1 𝛾𝛾−1
• 𝑇𝑇ℎ 𝑉𝑉𝐵𝐵 = 𝑇𝑇𝑐𝑐 𝑉𝑉𝐶𝐶 , 𝑇𝑇ℎ 𝑉𝑉𝐴𝐴 = 𝑇𝑇𝑐𝑐 𝑉𝑉𝐷𝐷
𝑉𝑉𝐵𝐵 𝛾𝛾−1 𝑉𝑉𝐶𝐶 𝛾𝛾−1 𝑉𝑉𝐵𝐵 𝑉𝑉𝐶𝐶
• Hence, = ⇒ =
𝑉𝑉𝐴𝐴 𝑉𝑉𝐷𝐷 𝑉𝑉𝐴𝐴 𝑉𝑉𝐷𝐷
|𝑄𝑄𝑐𝑐 | 𝑇𝑇𝑐𝑐 |𝑄𝑄𝑐𝑐 | 𝑇𝑇𝑐𝑐
• Hence, = ⇒ 𝑒𝑒𝐶𝐶 = 1 − =1−
|𝑄𝑄ℎ | 𝑇𝑇ℎ |𝑄𝑄ℎ | 𝑇𝑇ℎ
Carnot Efficiency
• All Carnot engines working between the same
two reservoirs have the same efficiency. This
efficiency, called the Carnot efficiency.
• Thus, Carnot efficiency must be independent
of the working substance of the engine.
• Hence it must depend only on the
temperatures of the reservoirs.
|𝑄𝑄𝑐𝑐 | 𝑇𝑇𝑐𝑐
• Carnot Efficiency: 𝑒𝑒𝐶𝐶 = 1 − =1−
|𝑄𝑄ℎ | 𝑇𝑇ℎ
Quiz
• Two moles of an ideal gas are carried around the
thermodynamic path ABCDA in the figure. Here
𝑇𝑇𝐷𝐷 = 150 K, 𝑇𝑇𝐴𝐴𝐴𝐴 = 300 K, 𝑇𝑇𝐵𝐵 = 600 K, and
𝑝𝑝𝐴𝐴 = 2.00 × 104 Pa, while 𝑝𝑝𝐷𝐷 = 1.00 × 104 Pa.
The volume 𝑉𝑉𝐴𝐴 = 0.250 𝑚𝑚3 , while 𝑉𝑉𝐵𝐵 = 0.50 𝑚𝑚3 .
• Find the work done, the heat lost or absorbed,
and the internal energy of the system for the
thermodynamic paths: (a) AB, (b) BC, (c) CD and
(d) DA.
• (e) Find the efficiency of that engine. [1×5=5]
(Given: 𝐶𝐶𝑝𝑝 = 20.8 J/mol.K, R = 8.314 J/mol.K)
Quiz
• Along 𝐴𝐴 → 𝐵𝐵 work done:
𝑊𝑊𝐴𝐴𝐴𝐴 = −𝑃𝑃Δ𝑉𝑉 = −𝑝𝑝𝐴𝐴 𝑉𝑉𝐵𝐵 − 𝑉𝑉𝐴𝐴
104 𝑁𝑁
= − 2.00 × 0.5 − 0.25 𝑚𝑚3
𝑚𝑚2
3
= −5.00 × 10 J
• Heat transferred:
𝑄𝑄𝐴𝐴𝐴𝐴 = 𝑛𝑛𝐶𝐶𝑃𝑃 Δ𝑇𝑇 = 𝑛𝑛𝐶𝐶𝑃𝑃 𝑇𝑇𝐵𝐵 − 𝑇𝑇𝐴𝐴𝐴𝐴
= (2.00 𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚)(20.8𝐽𝐽/𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚. 𝐾𝐾)(600𝐾𝐾 − 300𝐾𝐾)
= 1.248 × 104 J
• Change in internal energy: Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄𝐴𝐴𝐴𝐴 + 𝑊𝑊𝐴𝐴𝐴𝐴 =
7.48 × 103 J
Quiz
• On path BC: Work done=0, isochoric process.
• Heat transferred:
𝑄𝑄𝐵𝐵𝐵𝐵 = 𝑛𝑛 𝐶𝐶𝑉𝑉 Δ𝑇𝑇 = 𝑛𝑛(𝐶𝐶𝑃𝑃 − 𝑅𝑅)(𝑇𝑇𝐴𝐴𝐴𝐴 − 𝑇𝑇𝐵𝐵 )
𝐽𝐽
= 2 𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚 20.8 − 8.314 300 − 600 𝐾𝐾
𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚.𝐾𝐾
= −7.4916 × 103 J
• Change of internal energy: Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄𝐵𝐵𝐵𝐵
= −7.4916 × 103 J
Quiz
• On path CD, work done: 𝑊𝑊𝐶𝐶𝐶𝐶4 = −𝑃𝑃Δ𝑉𝑉 =
10 𝑁𝑁
− 𝑝𝑝𝐷𝐷 (𝑉𝑉𝐴𝐴 − 𝑉𝑉𝐵𝐵 ) = − 1 × 2 0.25 − 0.50 𝑚𝑚3
𝑚𝑚
= 2.5 × 103 J
• Heat transferred: 𝑄𝑄𝐶𝐶𝐶𝐶 = 𝑛𝑛𝐶𝐶𝑃𝑃 Δ𝑇𝑇
= 𝑛𝑛𝐶𝐶𝑃𝑃 𝑇𝑇𝐷𝐷 − 𝑇𝑇𝐴𝐴𝐴𝐴
20.8𝐽𝐽
= 2.0𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚 150 − 300 𝐾𝐾
𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚.𝐾𝐾
= −6.24 × 103 J
• Change in internal energy: Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄𝐶𝐶𝐶𝐶 + 𝑊𝑊𝐶𝐶𝐶𝐶
= −3.74 × 103 J
Quiz
• On path DA: Work done=0, isochoric process.
• Heat transferred:
𝑄𝑄𝐷𝐷𝐷𝐷 = 𝑛𝑛𝐶𝐶𝑉𝑉 Δ𝑇𝑇 = 𝑛𝑛(𝐶𝐶𝑃𝑃 − 𝑅𝑅)(𝑇𝑇𝐴𝐴𝐴𝐴 − 𝑇𝑇𝐷𝐷 )
𝐽𝐽
= 2𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚 20.8 − 8.314 300 − 150 𝐾𝐾
𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚.𝐾𝐾
= 3.7458 × 103 J
• Change in internal energy: Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄𝐷𝐷𝐷𝐷
Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 3.7458 × 103 J
Quiz
• Net work done:
𝑊𝑊𝐴𝐴𝐴𝐴𝐴𝐴𝐴𝐴𝐴𝐴 = 𝑊𝑊𝐴𝐴𝐴𝐴 + 𝑊𝑊𝐵𝐵𝐵𝐵 + 𝑊𝑊𝐶𝐶𝐶𝐶 + 𝑊𝑊𝐷𝐷𝐷𝐷
= −5.0 + 0 + 2.5 + 0 × 103 J = −2.5 × 103 J
• Net heat added:
𝑄𝑄𝐴𝐴𝐴𝐴𝐴𝐴𝐴𝐴𝐴𝐴 = 𝑄𝑄𝐴𝐴𝐴𝐴 + 𝑄𝑄𝐵𝐵𝐵𝐵 + 𝑄𝑄𝐶𝐶𝐶𝐶 + 𝑄𝑄𝐷𝐷𝐷𝐷
= 1.248 × 104 𝐽𝐽 − 7.4916 × 103 𝐽𝐽
−6.24 × 103 𝐽𝐽 + 3.7458 × 103 𝐽𝐽
= 2.4942 × 103 J ≈ 2.5 × 103 J
• Efficiency: 𝜂𝜂 = 𝑊𝑊𝑛𝑛𝑛𝑛𝑛𝑛−𝑜𝑜𝑜𝑜𝑜𝑜 /𝑄𝑄𝑖𝑖𝑖𝑖
2.5×103
= 2.5 × 103 𝐽𝐽/(𝑄𝑄𝐴𝐴𝐴𝐴 + 𝑄𝑄𝐷𝐷𝐷𝐷 )= = 15.407%
1.62258×104
Reversible and Irreversible Processes
• Reversible Process: In a reversible process,
the system undergoing the process can be
returned to its initial conditions along the
same path on a PV diagram.
• Every point along a reversible path is an
equilibrium state.
• Irreversible Process: In an irreversible process,
a system can not be taken in the backward
direction along its path.
Nature of Reversible and Irreversible
Processes
• Irreversible Process: Processes involving
turbulence, friction or dissipative forces.
• Reversible Process: Processes running very
slowly, such that the system is always very nearly
in equilibrium state (quasi-static process).
• No dissipative force present.
• Energy transfer as heat can only occur between
objects at the same/nearby temperatures.
Second Law, Irreversibility and
Impossibility
• Second law states which processes do and
which do not occur.
• No irreversible process has ever been
observed to run backward.
• Irreversible processes are restricted by the
second law to run in the opposite direction.
• Statements of second law (Kelvin–Planck or
Clausius forms) are statements of
impossibility.
Second Law, Irreversibility and State
Variable
• We can form an equivalent statement of
second law by using the concept of
irreversibility.
• To do this we have to introduce a new state
variable called the entropy.
• Zeroth law -> Concept of temperature.
• First law -> Concept of internal energy.
• Both 𝑇𝑇 and 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 are state variables (like 𝑃𝑃, 𝑉𝑉)
Second Law, Irreversibility and
Entropy
• Second law -> Concept of entropy S.
• Entropy increases in an irreversible process.
• In a reversible process, entropy remains the
same.
• Hence, entropy never decreases (a statement
of impossibility).
The entropy of an isolated system either
increases or remains constant in any process, it
never decreases.
Concept of Entropy –Necessity
• In a closed/isolated system irreversible processes
do not violate conservation of energy.
• Changes in energy Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 , do not set the
direction of a process (whether can occur
backward i.e. is a reversible process, or can not
occur backward i.e. it is an irreversible process).
• Changes in entropy Δ𝑆𝑆 , dictates whether a
process will occur or not (e.g. an irreversible
process will occur only in the forward direction
only, not in the backward.)
Entropy and Conservation Law
• Entropy differs from energy in that entropy
does not obey a conservation law.
• Energy of a closed system is conserved; it
always remains constant.
• For irreversible processes, the entropy of a
closed system always increases Δ𝑆𝑆 > 0. For
reversible process Δ𝑆𝑆 = 0.
• The change in entropy is sometimes called
“the arrow of time.”
Entropy and the Arrow of Time
• Example: Popping up of a popcorn is
associated with the forward direction of time.
• Associated with it is an increase in entropy.
• The backward direction of time (a videotape
running backwards) would correspond to the
exploded popcorn reforming the original
kernel/corn.
• Backward process implies Δ𝑆𝑆 < 0, and never
happens.
Entropy -Definitions
• Two equivalent definitions of entropy:
• A. In terms of system’s temperature and the
energy the system gains or loses as heat:
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑄𝑄𝑟𝑟𝑟𝑟𝑟𝑟 /𝑇𝑇
• B. In terms of counting the ways in which the
atoms or molecules that make up the system
can be arranged:
𝑆𝑆 = 𝑘𝑘𝐵𝐵 ln(𝑤𝑤)
𝑤𝑤=multiplicity of arrangements
Calculation of Entropy
• The change in entropy Δ𝑆𝑆, for an irreversible
process that takes a system from an initial
state 𝑖𝑖 to a final state 𝑓𝑓
= the change Δ𝑆𝑆 in entropy for any reversible
process that takes the system between those
same two states.
• We can compute the latter (but not the
former).
Entropy – Thermodynamic Definition
• For a reversible process between initial state 𝑖𝑖
and final state 𝑓𝑓, the change in entropy
𝑓𝑓 𝑓𝑓
𝑑𝑑𝑄𝑄𝑟𝑟𝑟𝑟𝑟𝑟
Δ𝑆𝑆 = 𝑆𝑆𝑓𝑓 − 𝑆𝑆𝑖𝑖 = � 𝑑𝑑𝑑𝑑 = �
𝑖𝑖 𝑖𝑖 𝑇𝑇
• 𝑑𝑑𝑄𝑄𝑟𝑟𝑟𝑟𝑟𝑟 = infinitesimal energy transferred as
heat, to or from the system during the process
• 𝑇𝑇 = temperature during the infinitesimal
energy transfer of amount 𝑑𝑑𝑄𝑄𝑟𝑟𝑟𝑟𝑟𝑟 as heat
Entropy – Thermodynamic Definition
• Theorem: For any process, reversible or
irreversible:
𝑓𝑓 𝑓𝑓
𝑑𝑑𝑄𝑄𝑟𝑟𝑟𝑟𝑟𝑟
Δ𝑆𝑆 = 𝑆𝑆𝑓𝑓 − 𝑆𝑆𝑖𝑖 = � 𝑑𝑑𝑑𝑑 = �
𝑖𝑖 𝑖𝑖 𝑇𝑇
• Proof: Any reversible process can be thought of a
combination of small steps of isothermal and
adiabatic processes.
• During adiabatic part of the step 𝑑𝑑𝑑𝑑 = 0
• During isothermal part of the step, 𝑇𝑇 = const.
Entropy – Thermodynamic Definition
• 𝑎𝑎𝑎𝑎𝑎𝑎𝑎𝑎𝑎𝑎 is a Carnot cycle.
𝑃𝑃
• 𝑎𝑎𝑎𝑎, 𝑐𝑐𝑐𝑐 isotherms.
• 𝑏𝑏𝑏𝑏, 𝑑𝑑𝑑𝑑 adiabats.
• For a Carnot cycle:
|𝑄𝑄𝑐𝑐 | 𝑇𝑇𝑐𝑐
we had =
|𝑄𝑄ℎ | 𝑇𝑇ℎ
𝑄𝑄1 −𝑄𝑄2
• ⇒ = , (𝑇𝑇1 > 𝑇𝑇2 )
𝑇𝑇1 𝑇𝑇2
𝑄𝑄1 𝑄𝑄2 𝑑𝑑𝑄𝑄1 𝑑𝑑𝑄𝑄2 𝑉𝑉
• ⇒ + = 0, ⇒ + = 0.
𝑇𝑇1 𝑇𝑇2 𝑇𝑇1 𝑇𝑇2
Entropy – Thermodynamic Definition
• 𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒 is also Carnot cycle. 𝑃𝑃
• 𝑒𝑒𝑒𝑒, 𝑔𝑔𝑔 isotherms.
• 𝑓𝑓𝑓𝑓, ℎ𝑒𝑒 adiabats.
𝑄𝑄3 𝑄𝑄4
• ⇒ + = 0,
𝑇𝑇3 𝑇𝑇4
𝑑𝑑𝑄𝑄3 𝑑𝑑𝑄𝑄4
⇒ + =0
𝑇𝑇3 𝑇𝑇4
𝑉𝑉
• Hence over the complete cycle:
∑𝑖𝑖 𝑑𝑑𝑄𝑄𝑖𝑖 /𝑇𝑇𝑖𝑖 = 0 ⇒ ∮ 𝑑𝑑𝑑𝑑/𝑇𝑇 = 0 (reversible)
Entropy – A State Variable
• Entropy change in a reversible cycle:

� 𝑑𝑑𝑄𝑄𝑟𝑟𝑟𝑟𝑟𝑟 /𝑇𝑇 = � 𝑑𝑑𝑆𝑆 = 0

• Now, we approximated an arbitrary reversible


path as combination of small steps consisting of
isotherms + adiabats.
• Since, ∮ 𝑑𝑑𝑑𝑑 = 0, for any reversible closed path,
we can think of many closed paths between the
same two points.
Entropy – A State Variable
• We can go from the same initial state 𝑖𝑖 to the
final state 𝑓𝑓 by different paths.
• Any closed path may consist of:
one path + (− other path).
• Since entropy change is zero in the closed path:
∫ 𝑑𝑑𝑑𝑑 = Δ𝑆𝑆 is the same
for all paths connecting the same
points. For example,
• (Δ𝐶𝐶𝐶𝐶𝐶𝐶 +Δ𝐴𝐴𝐴𝐴 )𝑆𝑆 = (Δ𝐶𝐶𝐷𝐷𝐴𝐴 +Δ𝐴𝐴𝐴𝐴𝐴𝐴 )𝑆𝑆
⇒ Δ𝐴𝐴𝐴𝐴 𝑆𝑆 = Δ𝐴𝐴𝐵𝐵𝐶𝐶 𝑆𝑆
Entropy – A State Variable
• Hence, between two points, 𝑖𝑖 and 𝑓𝑓, Δ𝑆𝑆 = 𝑆𝑆𝑓𝑓 −
𝑓𝑓
𝑆𝑆𝑖𝑖 = ∫𝑖𝑖 𝑑𝑑𝑑𝑑 is independent of path taken.
• Hence, entropy must be a function of the state or
a state variable.
• Although, we proved it for reversible paths, this
property is independent of path and true always.
𝑓𝑓
• Hence, for any process Δ𝑆𝑆 = 𝑆𝑆𝑓𝑓 − 𝑆𝑆𝑖𝑖 = ∫𝑖𝑖 𝑑𝑑𝑑𝑑 =
𝑓𝑓 𝑑𝑑𝑄𝑄𝑟𝑟𝑟𝑟𝑟𝑟
∫𝑖𝑖 . Hence, proved.
𝑇𝑇
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Quiz
• Two moles of an ideal gas are carried around the
thermodynamic path ABCDA in the figure. Here
𝑇𝑇𝐷𝐷 = 150 K, 𝑇𝑇𝐴𝐴𝐴𝐴 = 300 K, 𝑇𝑇𝐵𝐵 = 600 K, and
𝑝𝑝𝐴𝐴 = 2.00 × 104 Pa, while 𝑝𝑝𝐷𝐷 = 1.00 × 104 Pa.
The volume 𝑉𝑉𝐴𝐴 = 0.250 𝑚𝑚3 , while 𝑉𝑉𝐵𝐵 = 0.50 𝑚𝑚3 .
• Find the work done, the heat lost or absorbed,
and the internal energy of the system for the
thermodynamic paths: (a) AB, (b) BC, (c) CD and
(d) DA.
• (e) Find the efficiency of that engine. [1×5=5]
(Given: 𝐶𝐶𝑝𝑝 = 20.8 J/mol.K, R = 8.314 J/mol.K)
Quiz
• Along 𝐴𝐴 → 𝐵𝐵 work done:
𝑊𝑊𝐴𝐴𝐴𝐴 = −𝑃𝑃Δ𝑉𝑉 = −𝑝𝑝𝐴𝐴 𝑉𝑉𝐵𝐵 − 𝑉𝑉𝐴𝐴
104 𝑁𝑁
= − 2.00 × 0.5 − 0.25 𝑚𝑚3
𝑚𝑚2
3
= −5.00 × 10 J
• Heat transferred:
𝑄𝑄𝐴𝐴𝐴𝐴 = 𝑛𝑛𝐶𝐶𝑃𝑃 Δ𝑇𝑇 = 𝑛𝑛𝐶𝐶𝑃𝑃 𝑇𝑇𝐵𝐵 − 𝑇𝑇𝐴𝐴𝐴𝐴
= (2.00 𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚)(20.8𝐽𝐽/𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚. 𝐾𝐾)(600𝐾𝐾 − 300𝐾𝐾)
= 1.248 × 104 J
• Change in internal energy: Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄𝐴𝐴𝐴𝐴 + 𝑊𝑊𝐴𝐴𝐴𝐴 =
7.48 × 103 J
Quiz
• On path BC: Work done=0, isochoric process.
• Heat transferred:
𝑄𝑄𝐵𝐵𝐵𝐵 = 𝑛𝑛 𝐶𝐶𝑉𝑉 Δ𝑇𝑇 = 𝑛𝑛(𝐶𝐶𝑃𝑃 − 𝑅𝑅)(𝑇𝑇𝐴𝐴𝐴𝐴 − 𝑇𝑇𝐵𝐵 )
𝐽𝐽
= 2 𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚 20.8 − 8.314 300 − 600 𝐾𝐾
𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚.𝐾𝐾
= −7.4916 × 103 J
• Change of internal energy: Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄𝐵𝐵𝐵𝐵
= −7.4916 × 103 J
Quiz
• On path CD, work done: 𝑊𝑊𝐶𝐶𝐶𝐶4 = −𝑃𝑃Δ𝑉𝑉 =
10 𝑁𝑁
− 𝑝𝑝𝐷𝐷 (𝑉𝑉𝐴𝐴 − 𝑉𝑉𝐵𝐵 ) = − 1 × 2 0.25 − 0.50 𝑚𝑚3
𝑚𝑚
= 2.5 × 103 J
• Heat transferred: 𝑄𝑄𝐶𝐶𝐶𝐶 = 𝑛𝑛𝐶𝐶𝑃𝑃 Δ𝑇𝑇
= 𝑛𝑛𝐶𝐶𝑃𝑃 𝑇𝑇𝐷𝐷 − 𝑇𝑇𝐴𝐴𝐴𝐴
20.8𝐽𝐽
= 2.0𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚 150 − 300 𝐾𝐾
𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚.𝐾𝐾
= −6.24 × 103 J
• Change in internal energy: Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄𝐶𝐶𝐶𝐶 + 𝑊𝑊𝐶𝐶𝐶𝐶
= −3.74 × 103 J
Quiz
• On path DA: Work done=0, isochoric process.
• Heat transferred:
𝑄𝑄𝐷𝐷𝐷𝐷 = 𝑛𝑛𝐶𝐶𝑉𝑉 Δ𝑇𝑇 = 𝑛𝑛(𝐶𝐶𝑃𝑃 − 𝑅𝑅)(𝑇𝑇𝐴𝐴𝐴𝐴 − 𝑇𝑇𝐷𝐷 )
𝐽𝐽
= 2𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚 20.8 − 8.314 300 − 150 𝐾𝐾
𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚.𝐾𝐾
= 3.7458 × 103 J
• Change in internal energy: Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄𝐷𝐷𝐷𝐷
Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 3.7458 × 103 J
Quiz
• Net work done:
𝑊𝑊𝐴𝐴𝐴𝐴𝐴𝐴𝐴𝐴𝐴𝐴 = 𝑊𝑊𝐴𝐴𝐴𝐴 + 𝑊𝑊𝐵𝐵𝐵𝐵 + 𝑊𝑊𝐶𝐶𝐶𝐶 + 𝑊𝑊𝐷𝐷𝐷𝐷
= −5.0 + 0 + 2.5 + 0 × 103 J = −2.5 × 103 J
• Net heat added:
𝑄𝑄𝐴𝐴𝐴𝐴𝐴𝐴𝐴𝐴𝐴𝐴 = 𝑄𝑄𝐴𝐴𝐴𝐴 + 𝑄𝑄𝐵𝐵𝐵𝐵 + 𝑄𝑄𝐶𝐶𝐶𝐶 + 𝑄𝑄𝐷𝐷𝐷𝐷
= 1.248 × 104 𝐽𝐽 − 7.4916 × 103 𝐽𝐽
−6.24 × 103 𝐽𝐽 + 3.7458 × 103 𝐽𝐽
= 2.4942 × 103 J ≈ 2.5 × 103 J
• Efficiency: 𝜂𝜂 = 𝑊𝑊𝑛𝑛𝑛𝑛𝑛𝑛−𝑜𝑜𝑜𝑜𝑜𝑜 /𝑄𝑄𝑖𝑖𝑖𝑖
2.5×103
= 2.5 × 103 𝐽𝐽/(𝑄𝑄𝐴𝐴𝐴𝐴 + 𝑄𝑄𝐷𝐷𝐷𝐷 )= = 15.407%
1.62258×104
Reversible and Irreversible Processes
• Reversible Process: In a reversible process,
the system undergoing the process can be
returned to its initial conditions along the
same path on a PV diagram.
• Every point along a reversible path is an
equilibrium state.
• Irreversible Process: In an irreversible process,
a system can not be taken in the backward
direction along its path.
Nature of Reversible and Irreversible
Processes
• Irreversible Process: Processes involving
turbulence, friction or dissipative forces.
• Reversible Process: Processes running very
slowly, such that the system is always very nearly
in equilibrium state (quasi-static process).
• No dissipative force present.
• Energy transfer as heat can only occur between
objects at the same/nearby temperatures.
Second Law, Irreversibility and
Impossibility
• Second law states which processes do and
which do not occur.
• No irreversible process has ever been
observed to run backward.
• Irreversible processes are restricted by the
second law to run in the opposite direction.
• Statements of second law (Kelvin–Planck or
Clausius forms) are statements of
impossibility.
Second Law, Irreversibility and State
Variable
• We can form an equivalent statement of
second law by using the concept of
irreversibility.
• To do this we have to introduce a new state
variable called the entropy.
• Zeroth law -> Concept of temperature.
• First law -> Concept of internal energy.
• Both 𝑇𝑇 and 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 are state variables (like 𝑃𝑃, 𝑉𝑉)
Second Law, Irreversibility and
Entropy
• Second law -> Concept of entropy S.
• Entropy increases in an irreversible process.
• In a reversible process, entropy remains the
same.
• Hence, entropy never decreases (a statement
of impossibility).
The entropy of an isolated system either
increases or remains constant in any process, it
never decreases.
Concept of Entropy –Necessity
• In a closed/isolated system irreversible processes
do not violate conservation of energy.
• Changes in energy Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 , do not set the
direction of a process (whether can occur
backward i.e. is a reversible process, or can not
occur backward i.e. it is an irreversible process).
• Changes in entropy Δ𝑆𝑆 , dictates whether a
process will occur or not (e.g. an irreversible
process will occur only in the forward direction
only, not in the backward.)
Entropy and Conservation Law
• Entropy differs from energy in that entropy
does not obey a conservation law.
• Energy of a closed system is conserved; it
always remains constant.
• For irreversible processes, the entropy of a
closed system always increases Δ𝑆𝑆 > 0. For
reversible process Δ𝑆𝑆 = 0.
• The change in entropy is sometimes called
“the arrow of time.”
Entropy and the Arrow of Time
• Example: Popping up of a popcorn is
associated with the forward direction of time.
• Associated with it is an increase in entropy.
• The backward direction of time (a videotape
running backwards) would correspond to the
exploded popcorn reforming the original
kernel/corn.
• Backward process implies Δ𝑆𝑆 < 0, and never
happens.
Entropy -Definitions
• Two equivalent definitions of entropy:
• A. In terms of system’s temperature and the
energy the system gains or loses as heat:
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑄𝑄𝑟𝑟𝑟𝑟𝑟𝑟 /𝑇𝑇
• B. In terms of counting the ways in which the
atoms or molecules that make up the system
can be arranged:
𝑆𝑆 = 𝑘𝑘𝐵𝐵 ln(𝑤𝑤)
𝑤𝑤=multiplicity of arrangements
Calculation of Entropy
• The change in entropy Δ𝑆𝑆, for an irreversible
process that takes a system from an initial
state 𝑖𝑖 to a final state 𝑓𝑓
= the change Δ𝑆𝑆 in entropy for any reversible
process that takes the system between those
same two states.
• We can compute the latter (but not the
former).
Entropy – Thermodynamic Definition
• For a reversible process between initial state 𝑖𝑖
and final state 𝑓𝑓, the change in entropy
𝑓𝑓 𝑓𝑓
𝑑𝑑𝑄𝑄𝑟𝑟𝑟𝑟𝑟𝑟
Δ𝑆𝑆 = 𝑆𝑆𝑓𝑓 − 𝑆𝑆𝑖𝑖 = � 𝑑𝑑𝑑𝑑 = �
𝑖𝑖 𝑖𝑖 𝑇𝑇
• 𝑑𝑑𝑄𝑄𝑟𝑟𝑟𝑟𝑟𝑟 = infinitesimal energy transferred as
heat, to or from the system during the process
• 𝑇𝑇 = temperature during the infinitesimal
energy transfer of amount 𝑑𝑑𝑄𝑄𝑟𝑟𝑟𝑟𝑟𝑟 as heat
Entropy – Thermodynamic Definition
• Theorem: For any process, reversible or
irreversible:
𝑓𝑓 𝑓𝑓
𝑑𝑑𝑄𝑄𝑟𝑟𝑟𝑟𝑟𝑟
Δ𝑆𝑆 = 𝑆𝑆𝑓𝑓 − 𝑆𝑆𝑖𝑖 = � 𝑑𝑑𝑑𝑑 = �
𝑖𝑖 𝑖𝑖 𝑇𝑇
• Proof: Any reversible process can be thought of a
combination of small steps of isothermal and
adiabatic processes.
• During adiabatic part of the step 𝑑𝑑𝑑𝑑 = 0
• During isothermal part of the step, 𝑇𝑇 = const.
Entropy – Thermodynamic Definition
• 𝑎𝑎𝑎𝑎𝑎𝑎𝑎𝑎𝑎𝑎 is a Carnot cycle.
𝑃𝑃
• 𝑎𝑎𝑎𝑎, 𝑐𝑐𝑐𝑐 isotherms.
• 𝑏𝑏𝑏𝑏, 𝑑𝑑𝑑𝑑 adiabats.
• For a Carnot cycle:
|𝑄𝑄𝑐𝑐 | 𝑇𝑇𝑐𝑐
we had =
|𝑄𝑄ℎ | 𝑇𝑇ℎ
𝑄𝑄1 −𝑄𝑄2
• ⇒ = , (𝑇𝑇1 > 𝑇𝑇2 )
𝑇𝑇1 𝑇𝑇2
𝑄𝑄1 𝑄𝑄2 𝑑𝑑𝑄𝑄1 𝑑𝑑𝑄𝑄2 𝑉𝑉
• ⇒ + = 0, ⇒ + = 0.
𝑇𝑇1 𝑇𝑇2 𝑇𝑇1 𝑇𝑇2
Entropy – Thermodynamic Definition
• 𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒 is also Carnot cycle.𝑃𝑃
• 𝑒𝑒𝑒𝑒, 𝑔𝑔𝑔 isotherms.
• 𝑓𝑓𝑓𝑓, ℎ𝑒𝑒 adiabats.
𝑄𝑄3 𝑄𝑄4
• ⇒ + = 0,
𝑇𝑇3 𝑇𝑇4
𝑑𝑑𝑄𝑄3 𝑑𝑑𝑄𝑄4
⇒ + =0
𝑇𝑇3 𝑇𝑇4
𝑉𝑉
• Hence over the complete cycle:
∑𝑖𝑖 𝑑𝑑𝑄𝑄𝑖𝑖 /𝑇𝑇𝑖𝑖 = 0 ⇒ ∮ 𝑑𝑑𝑑𝑑/𝑇𝑇 = 0 (reversible)
Entropy – A State Variable
• Entropy change in a reversible cycle:

� 𝑑𝑑𝑄𝑄𝑟𝑟𝑟𝑟𝑟𝑟 /𝑇𝑇 = � 𝑑𝑑𝑆𝑆 = 0

• Now, we approximated an arbitrary reversible


path as combination of small steps consisting of
isotherms + adiabats.
• Since, ∮ 𝑑𝑑𝑑𝑑 = 0, for any reversible closed path,
we can think of many closed paths between the
same two points.
Entropy – A State Variable
• We can go from the same initial state 𝑖𝑖 to the
final state 𝑓𝑓 by different paths.
• Any closed path may consist of:
one path + (− other path).
• Since entropy change is zero in the closed path:
∫ 𝑑𝑑𝑑𝑑 = Δ𝑆𝑆 is the same
for all paths connecting the same
points. For example,
• (Δ𝐶𝐶𝐶𝐶𝐶𝐶 +Δ𝐴𝐴𝐴𝐴 )𝑆𝑆 = (Δ𝐶𝐶𝐷𝐷𝐴𝐴 +Δ𝐴𝐴𝐴𝐴𝐴𝐴 )𝑆𝑆
⇒ Δ𝐴𝐴𝐴𝐴 𝑆𝑆 = Δ𝐴𝐴𝐵𝐵𝐶𝐶 𝑆𝑆
Entropy – A State Variable
• Hence, between two points, 𝑖𝑖 and 𝑓𝑓, Δ𝑆𝑆 = 𝑆𝑆𝑓𝑓 −
𝑓𝑓
𝑆𝑆𝑖𝑖 = ∫𝑖𝑖 𝑑𝑑𝑑𝑑 is independent of path taken.
• Hence, entropy must be a function of the state or
a state variable.
• Although, we proved it for reversible paths, this
property is independent of path and true always.
𝑓𝑓
• Hence, for any process Δ𝑆𝑆 = 𝑆𝑆𝑓𝑓 − 𝑆𝑆𝑖𝑖 = ∫𝑖𝑖 𝑑𝑑𝑑𝑑 =
𝑓𝑓 𝑑𝑑𝑄𝑄𝑟𝑟𝑟𝑟𝑟𝑟
∫𝑖𝑖 . Hence, proved.
𝑇𝑇
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Quiz
• Two moles of an ideal gas are carried around the
thermodynamic path ABCDA in the figure. Here
𝑇𝑇𝐷𝐷 = 150 K, 𝑇𝑇𝐴𝐴𝐴𝐴 = 300 K, 𝑇𝑇𝐵𝐵 = 600 K, and
𝑝𝑝𝐴𝐴 = 2.00 × 104 Pa, while 𝑝𝑝𝐷𝐷 = 1.00 × 104 Pa.
The volume 𝑉𝑉𝐴𝐴 = 0.250 𝑚𝑚3 , while 𝑉𝑉𝐵𝐵 = 0.50 𝑚𝑚3 .
• Find the work done, the heat lost or absorbed,
and the internal energy of the system for the
thermodynamic paths: (a) AB, (b) BC, (c) CD and
(d) DA.
• (e) Find the efficiency of that engine. [1×5=5]
(Given: 𝐶𝐶𝑝𝑝 = 20.8 J/mol.K, R = 8.314 J/mol.K)
Quiz
• Along 𝐴𝐴 → 𝐵𝐵 work done:
𝑊𝑊𝐴𝐴𝐴𝐴 = −𝑃𝑃Δ𝑉𝑉 = −𝑝𝑝𝐴𝐴 𝑉𝑉𝐵𝐵 − 𝑉𝑉𝐴𝐴
104 𝑁𝑁
= − 2.00 × 0.5 − 0.25 𝑚𝑚3
𝑚𝑚2
3
= −5.00 × 10 J
• Heat transferred:
𝑄𝑄𝐴𝐴𝐴𝐴 = 𝑛𝑛𝐶𝐶𝑃𝑃 Δ𝑇𝑇 = 𝑛𝑛𝐶𝐶𝑃𝑃 𝑇𝑇𝐵𝐵 − 𝑇𝑇𝐴𝐴𝐴𝐴
= (2.00 𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚)(20.8𝐽𝐽/𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚. 𝐾𝐾)(600𝐾𝐾 − 300𝐾𝐾)
= 1.248 × 104 J
• Change in internal energy: Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄𝐴𝐴𝐴𝐴 + 𝑊𝑊𝐴𝐴𝐴𝐴 =
7.48 × 103 J
Quiz
• On path BC: Work done=0, isochoric process.
• Heat transferred:
𝑄𝑄𝐵𝐵𝐵𝐵 = 𝑛𝑛 𝐶𝐶𝑉𝑉 Δ𝑇𝑇 = 𝑛𝑛(𝐶𝐶𝑃𝑃 − 𝑅𝑅)(𝑇𝑇𝐴𝐴𝐴𝐴 − 𝑇𝑇𝐵𝐵 )
𝐽𝐽
= 2 𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚 20.8 − 8.314 300 − 600 𝐾𝐾
𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚.𝐾𝐾
= −7.4916 × 103 J
• Change of internal energy: Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄𝐵𝐵𝐵𝐵
= −7.4916 × 103 J
Quiz
• On path CD, work done: 𝑊𝑊𝐶𝐶𝐶𝐶4 = −𝑃𝑃Δ𝑉𝑉 =
10 𝑁𝑁
− 𝑝𝑝𝐷𝐷 (𝑉𝑉𝐴𝐴 − 𝑉𝑉𝐵𝐵 ) = − 1 × 2 0.25 − 0.50 𝑚𝑚3
𝑚𝑚
= 2.5 × 103 J
• Heat transferred: 𝑄𝑄𝐶𝐶𝐶𝐶 = 𝑛𝑛𝐶𝐶𝑃𝑃 Δ𝑇𝑇
= 𝑛𝑛𝐶𝐶𝑃𝑃 𝑇𝑇𝐷𝐷 − 𝑇𝑇𝐴𝐴𝐴𝐴
20.8𝐽𝐽
= 2.0𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚 150 − 300 𝐾𝐾
𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚.𝐾𝐾
= −6.24 × 103 J
• Change in internal energy: Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄𝐶𝐶𝐶𝐶 + 𝑊𝑊𝐶𝐶𝐶𝐶
= −3.74 × 103 J
Quiz
• On path DA: Work done=0, isochoric process.
• Heat transferred:
𝑄𝑄𝐷𝐷𝐷𝐷 = 𝑛𝑛𝐶𝐶𝑉𝑉 Δ𝑇𝑇 = 𝑛𝑛(𝐶𝐶𝑃𝑃 − 𝑅𝑅)(𝑇𝑇𝐴𝐴𝐴𝐴 − 𝑇𝑇𝐷𝐷 )
𝐽𝐽
= 2𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚 20.8 − 8.314 300 − 150 𝐾𝐾
𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚.𝐾𝐾
= 3.7458 × 103 J
• Change in internal energy: Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄𝐷𝐷𝐷𝐷
Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 3.7458 × 103 J
Quiz
• Net work done:
𝑊𝑊𝐴𝐴𝐴𝐴𝐴𝐴𝐴𝐴𝐴𝐴 = 𝑊𝑊𝐴𝐴𝐴𝐴 + 𝑊𝑊𝐵𝐵𝐵𝐵 + 𝑊𝑊𝐶𝐶𝐶𝐶 + 𝑊𝑊𝐷𝐷𝐷𝐷
= −5.0 + 0 + 2.5 + 0 × 103 J = −2.5 × 103 J
• Net heat added:
𝑄𝑄𝐴𝐴𝐴𝐴𝐴𝐴𝐴𝐴𝐴𝐴 = 𝑄𝑄𝐴𝐴𝐴𝐴 + 𝑄𝑄𝐵𝐵𝐵𝐵 + 𝑄𝑄𝐶𝐶𝐶𝐶 + 𝑄𝑄𝐷𝐷𝐷𝐷
= 1.248 × 104 𝐽𝐽 − 7.4916 × 103 𝐽𝐽
−6.24 × 103 𝐽𝐽 + 3.7458 × 103 𝐽𝐽
= 2.4942 × 103 J ≈ 2.5 × 103 J
• Efficiency: 𝜂𝜂 = 𝑊𝑊𝑛𝑛𝑛𝑛𝑛𝑛−𝑜𝑜𝑜𝑜𝑜𝑜 /𝑄𝑄𝑖𝑖𝑖𝑖
2.5×103
= 2.5 × 103 𝐽𝐽/(𝑄𝑄𝐴𝐴𝐴𝐴 + 𝑄𝑄𝐷𝐷𝐷𝐷 )= = 15.407%
1.62258×104
Reversible and Irreversible Processes
• Reversible Process: In a reversible process,
the system undergoing the process can be
returned to its initial conditions along the
same path on a PV diagram.
• Every point along a reversible path is an
equilibrium state.
• Irreversible Process: In an irreversible process,
a system can not be taken in the backward
direction along its path.
Nature of Reversible and Irreversible
Processes
• Irreversible Process: Processes involving
turbulence, friction or dissipative forces.
• Reversible Process: Processes running very
slowly, such that the system is always very nearly
in equilibrium state (quasi-static process).
• No dissipative force present.
• Energy transfer as heat can only occur between
objects at the same/nearby temperatures.
Second Law, Irreversibility and
Impossibility
• Second law states which processes do and
which do not occur.
• No irreversible process has ever been
observed to run backward.
• Irreversible processes are restricted by the
second law to run in the opposite direction.
• Statements of second law (Kelvin–Planck or
Clausius forms) are statements of
impossibility.
Second Law, Irreversibility and State
Variable
• We can form an equivalent statement of
second law by using the concept of
irreversibility.
• To do this we have to introduce a new state
variable called the entropy.
• Zeroth law -> Concept of temperature.
• First law -> Concept of internal energy.
• Both 𝑇𝑇 and 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 are state variables (like 𝑃𝑃, 𝑉𝑉)
Second Law, Irreversibility and
Entropy
• Second law -> Concept of entropy S.
• Entropy increases in an irreversible process.
• In a reversible process, entropy remains the
same.
• Hence, entropy never decreases (a statement
of impossibility).
The entropy of an isolated system either
increases or remains constant in any process, it
never decreases.
Concept of Entropy –Necessity
• In a closed/isolated system irreversible processes
do not violate conservation of energy.
• Changes in energy Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 , do not set the
direction of a process (whether can occur
backward i.e. is a reversible process, or can not
occur backward i.e. it is an irreversible process).
• Changes in entropy Δ𝑆𝑆 , dictates whether a
process will occur or not (e.g. an irreversible
process will occur only in the forward direction
only, not in the backward.)
Entropy and Conservation Law
• Entropy differs from energy in that entropy
does not obey a conservation law.
• Energy of a closed system is conserved; it
always remains constant.
• For irreversible processes, the entropy of a
closed system always increases Δ𝑆𝑆 > 0. For
reversible process Δ𝑆𝑆 = 0.
• The change in entropy is sometimes called
“the arrow of time.”
Entropy and the Arrow of Time
• Example: Popping up of a popcorn is
associated with the forward direction of time.
• Associated with it is an increase in entropy.
• The backward direction of time (a videotape
running backwards) would correspond to the
exploded popcorn reforming the original
kernel/corn.
• Backward process implies Δ𝑆𝑆 < 0, and never
happens.
Entropy -Definitions
• Two equivalent definitions of entropy:
• A. In terms of system’s temperature and the
energy the system gains or loses as heat:
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑄𝑄𝑟𝑟𝑟𝑟𝑟𝑟 /𝑇𝑇
• B. In terms of counting the ways in which the
atoms or molecules that make up the system
can be arranged:
𝑆𝑆 = 𝑘𝑘𝐵𝐵 ln(𝑤𝑤)
𝑤𝑤=multiplicity of arrangements
Calculation of Entropy
• The change in entropy Δ𝑆𝑆, for an irreversible
process that takes a system from an initial
state 𝑖𝑖 to a final state 𝑓𝑓
= the change Δ𝑆𝑆 in entropy for any reversible
process that takes the system between those
same two states.
• We can compute the latter (but not the
former).
Entropy – Thermodynamic Definition
• For a reversible process between initial state 𝑖𝑖
and final state 𝑓𝑓, the change in entropy
𝑓𝑓 𝑓𝑓
𝑑𝑑𝑄𝑄𝑟𝑟𝑟𝑟𝑟𝑟
Δ𝑆𝑆 = 𝑆𝑆𝑓𝑓 − 𝑆𝑆𝑖𝑖 = � 𝑑𝑑𝑑𝑑 = �
𝑖𝑖 𝑖𝑖 𝑇𝑇
• 𝑑𝑑𝑄𝑄𝑟𝑟𝑟𝑟𝑟𝑟 = infinitesimal energy transferred as
heat, to or from the system during the process
• 𝑇𝑇 = temperature during the infinitesimal
energy transfer of amount 𝑑𝑑𝑄𝑄𝑟𝑟𝑟𝑟𝑟𝑟 as heat
Entropy – Thermodynamic Definition
• Theorem: For any process, reversible or
irreversible:
𝑓𝑓 𝑓𝑓
𝑑𝑑𝑄𝑄𝑟𝑟𝑟𝑟𝑟𝑟
Δ𝑆𝑆 = 𝑆𝑆𝑓𝑓 − 𝑆𝑆𝑖𝑖 = � 𝑑𝑑𝑑𝑑 = �
𝑖𝑖 𝑖𝑖 𝑇𝑇
• Proof: Any reversible process can be thought of a
combination of small steps of isothermal and
adiabatic processes.
• During adiabatic part of the step 𝑑𝑑𝑑𝑑 = 0
• During isothermal part of the step, 𝑇𝑇 = const.
Entropy – Thermodynamic Definition
• 𝑎𝑎𝑎𝑎𝑎𝑎𝑎𝑎𝑎𝑎 is a Carnot cycle.
𝑃𝑃
• 𝑎𝑎𝑎𝑎, 𝑐𝑐𝑐𝑐 isotherms.
• 𝑏𝑏𝑏𝑏, 𝑑𝑑𝑑𝑑 adiabats.
• For a Carnot cycle:
|𝑄𝑄𝑐𝑐 | 𝑇𝑇𝑐𝑐
we had =
|𝑄𝑄ℎ | 𝑇𝑇ℎ
𝑄𝑄1 −𝑄𝑄2
• ⇒ = , (𝑇𝑇1 > 𝑇𝑇2 )
𝑇𝑇1 𝑇𝑇2
𝑄𝑄1 𝑄𝑄2 𝑑𝑑𝑄𝑄1 𝑑𝑑𝑄𝑄2 𝑉𝑉
• ⇒ + = 0, ⇒ + = 0.
𝑇𝑇1 𝑇𝑇2 𝑇𝑇1 𝑇𝑇2
Entropy – Thermodynamic Definition
• 𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒𝑒 is also Carnot cycle.𝑃𝑃
• 𝑒𝑒𝑒𝑒, 𝑔𝑔𝑔 isotherms.
• 𝑓𝑓𝑓𝑓, ℎ𝑒𝑒 adiabats.
𝑄𝑄3 𝑄𝑄4
• ⇒ + = 0,
𝑇𝑇3 𝑇𝑇4
𝑑𝑑𝑄𝑄3 𝑑𝑑𝑄𝑄4
⇒ + =0
𝑇𝑇3 𝑇𝑇4
𝑉𝑉
• Hence over the complete cycle:
∑𝑖𝑖 𝑑𝑑𝑄𝑄𝑖𝑖 /𝑇𝑇𝑖𝑖 = 0 ⇒ ∮ 𝑑𝑑𝑑𝑑/𝑇𝑇 = 0 (reversible)
Entropy – A State Variable
• Entropy change in a reversible cycle:

� 𝑑𝑑𝑄𝑄𝑟𝑟𝑟𝑟𝑟𝑟 /𝑇𝑇 = � 𝑑𝑑𝑆𝑆 = 0

• Now, we approximated an arbitrary reversible


path as combination of small steps consisting of
isotherms + adiabats.
• Since, ∮ 𝑑𝑑𝑑𝑑 = 0, for any reversible closed path,
we can think of many closed paths between the
same two points.
Entropy – A State Variable
• We can go from the same initial state 𝑖𝑖 to the
final state 𝑓𝑓 by different paths.
• Any closed path may consist of:
one path + (− other path).
• Since entropy change is zero in the closed path:
∫ 𝑑𝑑𝑑𝑑 = Δ𝑆𝑆 is the same
for all paths connecting the same
points. For example,
• (Δ𝐶𝐶𝐶𝐶𝐶𝐶 +Δ𝐴𝐴𝐴𝐴 )𝑆𝑆 = (Δ𝐶𝐶𝐷𝐷𝐴𝐴 +Δ𝐴𝐴𝐴𝐴𝐴𝐴 )𝑆𝑆
⇒ Δ𝐴𝐴𝐴𝐴 𝑆𝑆 = Δ𝐴𝐴𝐵𝐵𝐶𝐶 𝑆𝑆
Entropy – A State Variable
• Hence, between two points, 𝑖𝑖 and 𝑓𝑓, Δ𝑆𝑆 = 𝑆𝑆𝑓𝑓 −
𝑓𝑓
𝑆𝑆𝑖𝑖 = ∫𝑖𝑖 𝑑𝑑𝑑𝑑 is independent of path taken.
• Hence, entropy must be a function of the state or
a state variable.
• Although, we proved it for reversible paths, this
property is independent of path and true always.
𝑓𝑓
• Hence, for any process Δ𝑆𝑆 = 𝑆𝑆𝑓𝑓 − 𝑆𝑆𝑖𝑖 = ∫𝑖𝑖 𝑑𝑑𝑑𝑑 =
𝑓𝑓 𝑑𝑑𝑄𝑄𝑟𝑟𝑟𝑟𝑟𝑟
∫𝑖𝑖 . Hence, proved.
𝑇𝑇
Entropy Change for an Ideal Gas
• That entropy is indeed a state
function/variable can be deduced only by
experiment.
• However, we can prove it as a specific case for
an ideal gas.
• Example: Entropy change for an Ideal Gas:
• Consider an ideal gas is taken through a
reversible process.
Entropy Change for an Ideal Gas
• The process is made reversible by taking it
slowly in a series of small steps.
• The gas is in an equilibrium state at the end of
each step.
• 𝑑𝑑𝑑𝑑 = energy transferred as heat, to or from
the gas, at each step.
• 𝑑𝑑𝑑𝑑 = work done on the gas.
• From first law: 𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
Entropy Change for an Ideal Gas
• But 𝑑𝑑𝑑𝑑 = −𝑃𝑃𝑃𝑃𝑃𝑃, and for ideal gas 𝑑𝑑𝑑𝑑𝑖𝑖𝑖𝑖𝑖𝑖 =
𝑛𝑛𝐶𝐶𝑉𝑉 𝑑𝑑𝑑𝑑
𝑑𝑑𝑑𝑑
• ⇒ 𝑑𝑑𝑑𝑑 = 𝑃𝑃𝑃𝑃𝑃𝑃 + 𝑛𝑛𝐶𝐶𝑉𝑉 𝑑𝑑𝑑𝑑 = 𝑛𝑛𝑛𝑛𝑛𝑛 + 𝑛𝑛𝐶𝐶𝑉𝑉 𝑑𝑑𝑑𝑑
𝑉𝑉
𝑑𝑑𝑑𝑑 𝑑𝑑𝑑𝑑 𝑑𝑑𝑑𝑑
• ⇒ = 𝑛𝑛𝑛𝑛 + 𝑛𝑛𝐶𝐶𝑉𝑉
𝑇𝑇 𝑉𝑉 𝑇𝑇
• Integrating each term between an arbitrary
initial state 𝑖𝑖 and an arbitrary final state 𝑓𝑓:
𝑓𝑓 𝑑𝑑𝑑𝑑 𝑓𝑓 𝑑𝑑𝑑𝑑 𝑓𝑓 𝑑𝑑𝑑𝑑
• ∫𝑖𝑖 𝑇𝑇 = 𝑛𝑛𝑛𝑛 ∫𝑖𝑖 + 𝑛𝑛𝐶𝐶𝑉𝑉 ∫𝑖𝑖
𝑉𝑉 𝑇𝑇
Entropy Change for an Ideal Gas
𝑉𝑉𝑓𝑓 𝑇𝑇𝑓𝑓
• Δ𝑆𝑆 = 𝑆𝑆𝑓𝑓 − 𝑆𝑆𝑖𝑖 = 𝑛𝑛𝑛𝑛 ln + 𝑛𝑛𝐶𝐶𝑉𝑉 ln
𝑉𝑉𝑖𝑖 𝑇𝑇𝑖𝑖
• We did not specify a particular reversible process
during integration.
• The results of integration must hold for all
reversible processes that take the gas from state 𝑖𝑖
to state 𝑓𝑓.
• Δ𝑆𝑆 depends on the properties of the initial and
final states i.e. on (𝑉𝑉𝑖𝑖 , 𝑇𝑇𝑖𝑖 ) and (𝑉𝑉𝑓𝑓 , 𝑇𝑇𝑓𝑓 ).
• 𝛥𝛥𝛥𝛥 does not depend on the path taken. (Proved)
Entropy Change
• Example-2: Change in Entropy in an Adiabatic
Free Expansion of an Ideal Gas: Suppose 1.0
mole of nitrogen gas is confined to the left
side of the container. We open the stopcock,
and the volume of the gas doubles. What is
the change in entropy of the gas for this
irreversible process?
(Treat the gas as ideal.)
Entropy Change
• Solution: The process is a free expansion of an
ideal gas.
• A. We can determine the entropy change for the
irreversible process (free expansion) by
calculating it for a reversible process, between
the same states.
• B. The temperature of the ideal gas does not
change in the free expansion. Equivalent the
reversible process should be an isothermal
expansion.
Entropy Change
• From first law: 𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 − 𝑃𝑃𝑃𝑃𝑃𝑃
• For free expansion, 𝑇𝑇 =const.⇒ 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 =const.
• We consider an equivalent isothermal process.
𝑓𝑓 𝑑𝑑𝑑𝑑 𝑉𝑉𝑓𝑓
• 𝑑𝑑𝑑𝑑 = 𝑃𝑃𝑃𝑃𝑃𝑃 ⇒ 𝑄𝑄 = ∫𝑖𝑖 𝑛𝑛𝑛𝑛 𝑇𝑇 = 𝑛𝑛𝑛𝑛𝑛𝑛 ln
𝑉𝑉 𝑉𝑉𝑖𝑖
• Change in entropy in an equivalent reversible
isothermal process:
𝑉𝑉𝑓𝑓
Δ𝑆𝑆 = 𝑆𝑆𝑓𝑓 − 𝑆𝑆𝑖𝑖 = 𝑛𝑛𝑛𝑛 ln
𝑉𝑉𝑖𝑖
Entropy Change
• 𝑛𝑛 = the number of moles of gas present
• Δ𝑆𝑆 = 𝑛𝑛𝑛𝑛 ln 2 = (1.00 mol)(8.314
J/molK)(ln 2) = +5.76 J/K
• Hence, the entropy change for the free
expansion (and for all other processes that
connect the same initial and final states)
= 5.76 J/K
• Note: Δ𝑆𝑆 is positive, the entropy increases in
this irreversible process.
Entropy Change
• Example-3: Entropy Change in a Calorimetric
Processes: A substance of mass 𝑚𝑚1 , specific heat
𝑐𝑐1 , and initial temperature 𝑇𝑇𝑐𝑐 is placed in thermal
contact with a second substance of mass 𝑚𝑚2 ,
specific heat 𝑐𝑐2 , and initial temperature 𝑇𝑇ℎ > 𝑇𝑇𝑐𝑐 .
• The two substances are contained in a
calorimeter so that no energy is lost to the
surroundings. The system of the two substances
is allowed to reach thermal equilibrium.
• What is the total entropy change for the system?
Entropy Change
• Solution: Since no heat loss occurs to the
surroundings, 𝑄𝑄𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐 = −𝑄𝑄ℎ𝑜𝑜𝑜𝑜
• 𝑚𝑚1 𝑐𝑐1 Δ𝑇𝑇𝑐𝑐 = −𝑚𝑚2 𝑐𝑐2 Δ𝑇𝑇ℎ ,
• But, Δ𝑇𝑇𝑐𝑐 = 𝑇𝑇𝑓𝑓 − 𝑇𝑇𝑐𝑐 , Δ𝑇𝑇ℎ = 𝑇𝑇𝑓𝑓 − 𝑇𝑇ℎ , where
𝑇𝑇ℎ > 𝑇𝑇𝑓𝑓 > 𝑇𝑇𝑐𝑐 .
• Hot substance at the initial temperature 𝑇𝑇ℎ is
slowly cooled to the temperature 𝑇𝑇𝑓𝑓 .
• 𝑚𝑚1 𝑐𝑐1 𝑇𝑇𝑓𝑓 − 𝑇𝑇𝑐𝑐 = −𝑚𝑚2 𝑐𝑐2 (𝑇𝑇𝑓𝑓 − 𝑇𝑇ℎ )
• ⇒ 𝑇𝑇𝑓𝑓 = (𝑚𝑚1 𝑐𝑐1 𝑇𝑇𝑐𝑐 + 𝑚𝑚2 𝑐𝑐2 𝑇𝑇ℎ )/(𝑚𝑚1 𝑐𝑐1 + 𝑚𝑚2 𝑐𝑐2 )
Entropy Change
• Solution (contd.):
• The actual process is irreversible.
• The system goes through a series of non-
equilibrium states.
• However we can think of an equivalent process in
which system remains in contact with a series of
temperature reservoirs differing infinitesimally in
temperature.
• The first reservoir being at 𝑇𝑇ℎ and the last being
at 𝑇𝑇𝑓𝑓 . Then we can think of an equivalent
reversible process.
Entropy Change
• Solution (contd.):
• The change in entropy is:
𝑓𝑓 𝑑𝑑𝑑𝑑 𝑓𝑓 𝑚𝑚1 𝑐𝑐1 𝑑𝑑𝑑𝑑 𝑓𝑓 𝑚𝑚2 𝑐𝑐2 𝑑𝑑𝑑𝑑
Δ𝑆𝑆 = ∫𝑖𝑖 𝑇𝑇 = ∫𝑖𝑖 � + ∫𝑖𝑖 �
𝑇𝑇 𝑐𝑐 𝑇𝑇 ℎ
𝑇𝑇𝑓𝑓 𝑑𝑑𝑑𝑑 𝑇𝑇𝑓𝑓 𝑑𝑑𝑑𝑑
• ⇒ Δ𝑆𝑆 = 𝑚𝑚1 𝑐𝑐1 ∫𝑇𝑇 + 𝑚𝑚2 𝑐𝑐2 ∫𝑇𝑇
𝑐𝑐 𝑇𝑇 ℎ 𝑇𝑇
𝑇𝑇𝑓𝑓 𝑇𝑇𝑓𝑓
• ⇒ Δ𝑆𝑆 = 𝑚𝑚1 𝑐𝑐1 ln + 𝑚𝑚2 𝑐𝑐2 ln
𝑇𝑇𝑐𝑐 𝑇𝑇ℎ
Entropy Change
• Solution (contd.): Here, 𝑇𝑇𝑓𝑓 = (𝑚𝑚1 𝑐𝑐1 𝑇𝑇𝑐𝑐 + 𝑚𝑚2 𝑐𝑐2 𝑇𝑇ℎ )/(𝑚𝑚1 𝑐𝑐1 +
𝑚𝑚2 𝑐𝑐2 ).
• Also,
𝑇𝑇𝑓𝑓
ln = ln 𝑚𝑚1 𝑐𝑐1 + 𝑚𝑚2 𝑐𝑐2 𝑇𝑇ℎ /𝑇𝑇𝑐𝑐 − ln(𝑚𝑚1 𝑐𝑐1 + 𝑚𝑚2 𝑐𝑐2 ) > 0
𝑇𝑇𝑐𝑐
𝑇𝑇𝑓𝑓
ln = ln 𝑚𝑚2 𝑐𝑐2 + 𝑚𝑚1 𝑐𝑐1 𝑇𝑇𝑐𝑐 /𝑇𝑇ℎ − ln(𝑚𝑚1 𝑐𝑐1 + 𝑚𝑚2 𝑐𝑐2 ) < 0
𝑇𝑇ℎ
𝑇𝑇𝑓𝑓 𝑇𝑇𝑓𝑓
• Hence, 𝑚𝑚1 𝑐𝑐1 ln > 𝑚𝑚2 𝑐𝑐2 ln , ⇒ Δ𝑆𝑆 > 0. Entropy
𝑇𝑇𝑐𝑐 𝑇𝑇ℎ
increases.
• Note: Here we assumed that there no mixing of the substances. If
mixing occurs, entropy change would be higher.
Entropy Change
• Example-4: Suppose that 1.00 kg of water at
0.00°C is mixed with an equal mass of water at
100°C. After equilibrium is reached, the
mixture has a uniform temperature of 50.0°C.
• What is the change in entropy of the system?
• Solution: Here, 𝑚𝑚1 = 𝑚𝑚2 = 1.00 kg,
𝑐𝑐1 = 𝑐𝑐2 = 4186 J/kg.K, 𝑇𝑇1 = 273 K, 𝑇𝑇2 = 373
K,
• 𝑇𝑇𝑓𝑓 = 323 K (since 𝑚𝑚1 = 𝑚𝑚2 , 𝑐𝑐1 = 𝑐𝑐2 ).
Entropy Change
• Solution (contd.):
𝑇𝑇𝑓𝑓 𝑇𝑇𝑓𝑓
Δ𝑆𝑆 = 𝑚𝑚1 𝑐𝑐1 ln + 𝑚𝑚2 𝑐𝑐2 ln
𝑇𝑇1 𝑇𝑇2
323𝐾𝐾
• ⇒ Δ𝑆𝑆 = 1.00 𝑘𝑘𝑘𝑘 4186𝐽𝐽/𝑘𝑘𝑘𝑘. 𝐾𝐾 ln
273𝐾𝐾
323𝐾𝐾
• + 1.00 𝑘𝑘𝑘𝑘 4186𝐽𝐽/𝑘𝑘𝑘𝑘. 𝐾𝐾 ln
373𝐾𝐾
• = 704𝐽𝐽/𝐾𝐾 −602 𝐽𝐽/𝐾𝐾 = 102 𝐽𝐽/𝐾𝐾 >0.
Entropy- ST diagram
• Example- ST diagram: Because entropy is a
state function, thermodynamic processes can
be represented as ST, SV, or SP diagrams
instead of the PV diagrams we have used so
far. Make a sketch of the Carnot cycle on an ST
plot.
• Solution: The Carnot cycle consists of four
steps:
Entropy- ST diagram
• A. A reversible isothermal expansion
• B. Followed by a reversible adiabatic expansion,
• C. Then a reversible isothermal compression
• D. Followed by a reversible adiabatic compression.
Entropy- ST diagram
• During the isothermal expansion (1 to 2) heat
is absorbed reversibly so, S increases at
constant T.

• During the reversible adiabatic expansion (2 to


3) the temperature decreases while S is
constant.
Entropy- ST diagram
• During the isothermal compression (3 to 4)
heat is rejected reversibly, so S decreases at
constant T.

• During the reversible adiabatic compression (4


to 1) the temperature increases while S is
constant.
Entropy and Disorder
• In statistical mechanics, the behavior of a
substance is described in terms of the
statistical behavior of its constituent parts (i.e.
atoms and molecules).
• Isolated systems tend toward disorder and
that entropy is a measure of this disorder.
• To see this, we will consider a problem.
Entropy and Disorder
• Example: Consider the distribution of gas
molecules between the two halves of an
insulated box.
• The figure shows a box that contains six
identical (and thus indistinguishable)
molecules of a gas.
Entropy and Disorder
• Because the two halves have equal volumes,
a molecule has the same likelihood, or
probability, of being in either half.
• The following table shows
the seven possible
configurations of the
six molecules:
Entropy and Disorder
• Microstate: A microstate is a particular
configuration of the individual constituents of
the system.
• Macrostate: A macrostate is a description of
the conditions of the system from a
macroscopic point of view and makes use of
macroscopic variables such as pressure,
density, and temperature for gases.
Entropy and Disorder
• Each of these configurations may have a
number of microstates.
• For example: A. In configuration I, all six
molecules are in the left half of the box
(𝑛𝑛1 = 6) and none are in the right half
(𝑛𝑛2 = 0). There is only one microstate.
• B. In general, since 𝑁𝑁 = 6, we can select the
first molecule in six independent ways; that is,
we can pick any one of the six molecules.
Entropy and Disorder
• We can pick the second molecule in five ways, by
picking any one of the remaining five molecules;
and so on.
• The total number of ways in which we can select
all six molecules is the product of these
independent ways: 6 × 5 × 4 × 3 × 2 × 1 =
720 = 6!
• However, because the molecules are
indistinguishable, these 720 arrangements are
not all different.
Entropy and Disorder
• For example, if 𝑛𝑛1 = 4 and 𝑛𝑛2 = 2, the order in
which you put four molecules in one half of the
box does not matter.
• Because after we have put all four in, there is no
way that we can tell the order in which you did
so.
• The number of ways in which we can order the
four molecules is 4! = 24.
• The number of ways in which we can order the
remaining two molecules is 2! = 2.
Entropy and Disorder
• Hence the number of different arrangements
that lead to the (4, 2) split (of configuration
III) is 720/(4! 2!).
• Thus the number of microstates that
correspond to the given configuration i.e. the
multiplicity 𝑤𝑤 of that configuration is:
6! 720
𝑤𝑤𝐼𝐼𝐼𝐼𝐼𝐼 = = = 15
4! 2! 24 × 2
Entropy and Disorder
• C. In general, for 𝑁𝑁 indistinguishable
molecules distributed in two halves, with 𝑛𝑛1
and 𝑛𝑛2 split, we have
𝑁𝑁!
𝑤𝑤 = (multiplicity of configuration)
𝑛𝑛1 !𝑛𝑛2 !
Entropy and Disorder
• All 64 microstates would occur equally often.
• Thus the system will spend, on average, the same
amount of time in each of the 64 microstates.
• Because all microstates are equally probable, the
configurations are not all equally probable.
• Configuration IV, with 20 microstates, is the most
probable configuration, with a probability of
20
= 0.313.
64
Entropy and Disorder
• The system is in configuration IV 31.3% of the
time.
• Configurations I and VII, in which all the
molecules are in one half of the box, are the least
1
probable, each with a probability of = 0.016
64
or 1.6%.
• The most probable configuration is the one in
which the molecules are evenly divided between
the two halves of the box.
Entropy and Disorder
• We expect this at thermal equilibrium.
• Hence, there is a relationship between
probability (or multiplicity) and entropy.
• In 1877, Austrian physicist Ludwig Boltzmann
gave a relationship between them:
𝑆𝑆 = 𝑘𝑘𝐵𝐵 ln(𝑤𝑤)
𝑤𝑤=multiplicity of arrangements
Entropy and Disorder
Entropy and Disorder
• We see the corresponding entropies of the
configurations are:
Entropy and Disorder
• Example-4: Entropy in Free Expansion: Consider
free expansion of an ideal gas.
• The gas is initially confined to a volume 𝑉𝑉𝑖𝑖
• When the partition separating 𝑉𝑉𝑖𝑖 from a larger
container is removed, the molecules eventually
are distributed throughout the greater volume
𝑉𝑉𝑓𝑓 .
Entropy and Disorder
• The instant after the partition is removed, all
the molecules are in the initial volume.
• Each molecule occupies some microscopic
volume 𝑉𝑉𝑚𝑚 .
• The total number of possible locations of a
single molecule, in a macroscopic initial
volume 𝑉𝑉𝑖𝑖 is: 𝑤𝑤𝑖𝑖 = 𝑉𝑉𝑖𝑖 /𝑉𝑉𝑚𝑚
• To see this, divide the volume in small boxes
of volume 𝑉𝑉𝑚𝑚 : 𝑉𝑉𝑚𝑚
Entropy and Disorder
• 𝑤𝑤𝑖𝑖 represents the number of ways that the
molecule can be placed in the volume.
• 𝑤𝑤𝑖𝑖 = the number of microstates, which is
equivalent to the number of available
locations.
• Assume that the probabilities of a molecule
occupying any of these locations are equal.
Entropy and Disorder
• As more molecules are added to the system, the
number of possible ways that the molecules can
be positioned in the volume multiplies.
• For two molecules, for every possible placement
of the first, all possible placements of the second
are available.
• There are 𝑤𝑤1 ways of locating the first molecule.
• For each of these, there are 𝑤𝑤2 ways of locating
the second molecule.
• The total number of ways of locating the two
molecules is 𝑤𝑤1 𝑤𝑤2 .
Entropy and Disorder
• Neglecting the very small probability of having
two molecules occupy the same location:
• Each molecule may go into any of the 𝑉𝑉𝑖𝑖 /𝑉𝑉𝑚𝑚
number of locations.
• The number of ways of locating 𝑁𝑁 molecules in
𝑉𝑉𝑖𝑖 𝑁𝑁
the volume 𝑉𝑉𝑖𝑖 is 𝑤𝑤𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑤𝑤𝑖𝑖𝑁𝑁 =
𝑉𝑉𝑚𝑚
• The number of ways of locating 𝑁𝑁 molecules in
𝑉𝑉𝑓𝑓 𝑁𝑁
the volume 𝑉𝑉𝑓𝑓 is 𝑤𝑤𝑓𝑓𝑓𝑓𝑓𝑓 = 𝑤𝑤𝑓𝑓𝑁𝑁 =
𝑉𝑉𝑚𝑚
Entropy and Disorder
• The ratio of the number of ways of placing the
molecules in the volume of the initial and final
configurations:
𝑁𝑁
𝑤𝑤𝑓𝑓𝑓𝑓𝑓𝑓 𝑉𝑉𝑓𝑓 /𝑉𝑉𝑚𝑚 𝑉𝑉𝑓𝑓 𝑁𝑁
• = 𝑁𝑁 =
𝑤𝑤𝑖𝑖𝑖𝑖𝑖𝑖 𝑉𝑉𝑖𝑖 /𝑉𝑉𝑚𝑚 𝑉𝑉𝑖𝑖
• Taking natural logarithm and multiplying by
𝑤𝑤𝑓𝑓𝑓𝑓𝑓𝑓 𝑉𝑉𝑓𝑓
𝑘𝑘𝐵𝐵 , we get 𝑘𝑘𝐵𝐵 ln = 𝑛𝑛𝑁𝑁𝐴𝐴 𝑘𝑘𝐵𝐵 ln
𝑤𝑤𝑖𝑖𝑖𝑖𝑖𝑖 𝑉𝑉𝑖𝑖
using 𝑁𝑁 = 𝑛𝑛𝑁𝑁𝐴𝐴 .
Entropy and Disorder
𝑤𝑤𝑓𝑓𝑓𝑓𝑓𝑓 𝑉𝑉𝑓𝑓
• Hence, 𝑘𝑘𝐵𝐵 ln = 𝑛𝑛𝑛𝑛 ln
𝑤𝑤𝑖𝑖𝑖𝑖𝑖𝑖 𝑉𝑉𝑖𝑖
• But we had, for adiabatic free expansion,
𝑉𝑉𝑓𝑓
Δ𝑆𝑆 = 𝑆𝑆𝑓𝑓 − 𝑆𝑆𝑖𝑖 = 𝑛𝑛𝑛𝑛 ln
𝑉𝑉𝑖𝑖
• Hence we can identify:
𝑆𝑆 = 𝑘𝑘𝐵𝐵 ln(𝑤𝑤)
• Entropy is a measure of disorder.
Entropy and Disorder
• Example-6: Suppose that there are 100
indistinguishable molecules in a box.
• How many microstates are associated with the
configuration 𝑛𝑛1 = 50 and 𝑛𝑛2 = 50, and with
the configuration 𝑛𝑛1 = 100 and 𝑛𝑛2 = 0?
• Interpret the results in terms of the relative
probabilities of the two configurations.
Entropy and Disorder
• Solution: The multiplicity 𝑤𝑤 of a configuration
of indistinguishable molecules in a closed box
= the number of independent microstates
with that configuration.
• For the (𝑛𝑛1 , 𝑛𝑛2 ) configuration (50, 50)
• 𝑤𝑤 = 𝑁𝑁!/(𝑛𝑛1 ! 𝑛𝑛2 !)=100!/(50! 50!)
9.33×10157
• 𝑤𝑤 = = 1.01 × 1029
(3.04×1064 )(3.04×1064 )
Entropy and Disorder
• For the (𝑛𝑛1 , 𝑛𝑛2 ) configuration (100, 0)
100!
• 𝑤𝑤 = 𝑁𝑁!/(𝑛𝑛1 ! 𝑛𝑛2 !)= =1
100!0!
• Thus, a 50 – 50 distribution is more likely
than a 100 – 0 distribution by the enormous
factor of about 1.01 × 1029 .
• How large is this number?
Entropy and Disorder
• If we count, at one per nanosecond, the
number of microstates that correspond to the
50 – 50 distribution, it would take you about
3 × 1012 years, which is about 200 times
longer than the age of the universe.
• 100 molecules used in this sample problem is
a very small number. Imagine what these
calculated probabilities would be like for a
mole of molecules, say about 𝑁𝑁 = 1024 !
Entropy and Disorder
• Stirling’s Approximation:
ln 𝑁𝑁! ≈ 𝑁𝑁(ln 𝑁𝑁) − 𝑁𝑁
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Review
• Ideal Gas Law:
𝑃𝑃𝑃𝑃 = 𝑛𝑛𝑛𝑛𝑛𝑛 = 𝑁𝑁𝑘𝑘𝐵𝐵 𝑇𝑇 = 𝑛𝑛𝑁𝑁𝐴𝐴 𝑘𝑘𝐵𝐵 𝑇𝑇
• Mechanical equivalent of heat:
1cal=4.186 J
• Work done on a gas in a piston-cylinder system:
𝑑𝑑𝑑𝑑 = −𝑃𝑃𝑃𝑃𝑃𝑃
• Gas compressed:
𝒅𝒅𝒅𝒅 < 𝟎𝟎, 𝒅𝒅𝒅𝒅 = −𝑷𝑷 𝒅𝒅𝒅𝒅 > 𝟎𝟎, Work done on gas >0
• Gas expands:
𝒅𝒅𝒅𝒅 > 𝟎𝟎, 𝒅𝒅𝒅𝒅 = −𝑷𝑷 𝒅𝒅𝒅𝒅 < 𝟎𝟎, Work done on gas < 0
Review
• Total work done on the gas, for volume from
𝑉𝑉𝑖𝑖 to 𝑉𝑉𝑓𝑓 is
𝑉𝑉𝑓𝑓
𝑊𝑊 = − � 𝑃𝑃(𝑉𝑉) 𝑑𝑑𝑑𝑑
𝑉𝑉𝑖𝑖
• First law of thermodynamics:
Δ 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖 + 𝑊𝑊𝑜𝑜𝑜𝑜
• In the infinitesimal form:
𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑑𝑑𝑄𝑄𝑖𝑖𝑖𝑖 + 𝑑𝑑𝑊𝑊𝑜𝑜𝑜𝑜
Review
• Isolated System: 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝑖𝑖 = 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,𝑓𝑓
• Cyclic Process: Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 0, 𝑄𝑄 = −𝑊𝑊 (cyclic)
• Adiabatic Process: Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑊𝑊
(a) Thermal insulation or (b) Rapid process
• Adiabatic compression: Work done on the system
𝑊𝑊 > 0 ⇒ Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 > 0, ⇒ Δ𝑇𝑇 > 0, temperature of
the system increases.
• Adiabatic expansion: Work done on the system
𝑊𝑊 < 0 ⇒ Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 < 0, ⇒ Δ𝑇𝑇 < 0, temperature of
the system decreases.
Review
• Adiabatic Free Expansion:
Q = W = 0 ⇒ Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 0
• Isobaric Process: 𝑃𝑃 = 𝑃𝑃𝑎𝑎𝑎𝑎𝑎𝑎 + 𝑀𝑀𝑀𝑀/𝐴𝐴 = 𝑃𝑃𝑔𝑔𝑔𝑔𝑔𝑔
𝑊𝑊𝑜𝑜𝑜𝑜 = −𝑃𝑃(𝑉𝑉𝑓𝑓 − 𝑉𝑉𝑖𝑖 ) 𝑃𝑃 =const.
• Isochoric Process: 𝑉𝑉 = constant
Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄 + 0 = 𝑄𝑄 (isochoric process)
• Isothermal Process: 𝑇𝑇 = constant
Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 0 = 𝑄𝑄 + 𝑊𝑊 (Isothermal process, ideal gas)
Review
• Isothermal Expansion: 𝑄𝑄𝑖𝑖𝑖𝑖 > 0, heat flows
into the system ⇒ 𝑊𝑊𝑜𝑜𝑜𝑜 < 0 i.e. work on the
system< 0 i.e. system does work ⇒ Volume
expands.
• Isothermal Compression: 𝑄𝑄𝑖𝑖𝑖𝑖 < 0, heat flows
out of the gas ⇒ 𝑊𝑊𝑜𝑜𝑜𝑜 > 0 i.e. work on by the
system> 0 ⇒ Volume decreases.
Review
• Isothermal Expansion of an Ideal Gas:
𝑉𝑉𝑓𝑓
𝑉𝑉𝑓𝑓
𝑛𝑛𝑛𝑛𝑛𝑛
𝑊𝑊𝑜𝑜𝑜𝑜 = − � 𝑃𝑃 𝑑𝑑𝑑𝑑 = − � 𝑑𝑑𝑑𝑑
𝑉𝑉𝑖𝑖 𝑉𝑉
𝑉𝑉𝑖𝑖
𝑊𝑊𝑜𝑜𝑜𝑜 = 𝑛𝑛𝑛𝑛𝑛𝑛ln 𝑉𝑉𝑖𝑖 /𝑉𝑉𝑓𝑓 =𝑛𝑛𝑛𝑛𝑛𝑛ln 𝑃𝑃𝑓𝑓 /𝑃𝑃𝑖𝑖
Review
• Kinetic Expression of Pressure:
𝐹𝐹 2 𝑁𝑁 1 2
2 𝑁𝑁
𝑃𝑃 = = 𝑚𝑚 𝑣𝑣 = 𝐾𝐾. 𝐸𝐸.
𝐴𝐴 3 𝑉𝑉 2 3 𝑉𝑉
• Thermal interpretation of temperature:
• Temperature is a direct measure of average molecular kinetic
energy: 𝑃𝑃𝑃𝑃 = 𝑁𝑁𝑘𝑘𝐵𝐵 𝑇𝑇
1 2
3
𝑚𝑚𝑣𝑣 = 𝑘𝑘𝐵𝐵 𝑇𝑇
2 2
𝑁𝑁 3𝑁𝑁 3
• 𝐾𝐾𝑡𝑡𝑡𝑡𝑡𝑡,𝑡𝑡𝑡𝑡𝑡𝑡 = 𝑁𝑁𝐾𝐾. 𝐸𝐸. = 𝑚𝑚𝑣𝑣 2 = 𝑘𝑘 𝑇𝑇 = 𝑛𝑛𝑛𝑛𝑛𝑛
2 2 𝐵𝐵 2
1 𝑁𝑁
• RMS speed: 𝑣𝑣𝑟𝑟𝑟𝑟𝑟𝑟 = 𝑣𝑣 2 = ∑𝑖𝑖=1 𝑣𝑣𝑖𝑖2
𝑁𝑁
𝑣𝑣𝑟𝑟𝑟𝑟𝑟𝑟 = 3 𝑘𝑘𝐵𝐵 𝑇𝑇/𝑚𝑚 = 3𝑅𝑅𝑅𝑅/𝑀𝑀
Review
• 𝑄𝑄𝑖𝑖𝑖𝑖,𝑉𝑉 = 𝑛𝑛𝐶𝐶𝑉𝑉 Δ𝑇𝑇 (constant volume)
• 𝑄𝑄𝑖𝑖𝑖𝑖,𝑃𝑃 = 𝑛𝑛𝐶𝐶𝑃𝑃 Δ𝑇𝑇 (constant pressure)
• Constant Pressure Specific Heat:
𝑄𝑄𝑖𝑖𝑖𝑖 > 0, Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 > 0
• Work is done by the gas because of the
change in volume ( Δ𝑉𝑉 > 0), ⇒ 𝑊𝑊𝑜𝑜𝑜𝑜 < 0
• 𝑄𝑄𝑃𝑃 = 𝑛𝑛𝐶𝐶𝑃𝑃 Δ𝑇𝑇 = Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 − 𝑊𝑊𝑜𝑜𝑜𝑜
Review
• Constant Volume Specific Heat:
𝑄𝑄𝑖𝑖𝑖𝑖 > 0, Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 > 0,
• No work is done on/by the gas: 𝑊𝑊 = −∫ 𝑃𝑃𝑃𝑃𝑃𝑃 = 0
Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄 = 𝑛𝑛𝐶𝐶𝑉𝑉 Δ𝑇𝑇
𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑛𝑛𝐶𝐶𝑉𝑉 𝑇𝑇 (Internal energy of an ideal gas)
• Molar Specific Heat of an Ideal Gas:
• At constant volume, for an ideal gas,
Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄𝑉𝑉 = 𝑛𝑛𝐶𝐶𝑉𝑉 Δ𝑇𝑇
Review
• At constant pressure, for an ideal gas:
𝑄𝑄𝑃𝑃 = 𝑛𝑛𝐶𝐶𝑃𝑃 Δ𝑇𝑇 = Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 − 𝑊𝑊𝑜𝑜𝑜𝑜
• Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄𝑃𝑃 + 𝑊𝑊𝑜𝑜𝑜𝑜 = 𝑛𝑛𝐶𝐶𝑃𝑃 Δ𝑇𝑇 + −𝑃𝑃Δ𝑉𝑉
• For same Δ𝑇𝑇 , Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 is the same for ideal gas
𝑛𝑛𝐶𝐶𝑉𝑉 Δ𝑇𝑇 = 𝑛𝑛𝐶𝐶𝑃𝑃 Δ𝑇𝑇 − 𝑛𝑛𝑛𝑛Δ𝑇𝑇 Δ 𝑃𝑃𝑃𝑃 = 𝑃𝑃Δ𝑉𝑉

𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉 = 𝑅𝑅 for an ideal gas


Review
3
• Mono-atomic gas, Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝐾𝐾𝑡𝑡𝑡𝑡𝑡𝑡,𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡 = 𝑛𝑛𝑛𝑛𝑛𝑛
2
𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 3 3
𝐶𝐶𝑉𝑉 = = 𝑛𝑛𝑛𝑛𝑛𝑛/(𝑛𝑛𝑛𝑛) = 𝑅𝑅 (Mono-atomic gas)
𝑛𝑛𝑛𝑛 2 2
3 3 5
• 𝐶𝐶𝑃𝑃 /𝐶𝐶𝑉𝑉 = ( + 1)/( ) = = 1.67
2 2 3
• Diatomic Molecule:
1 5 5
𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑁𝑁 3 + 2 𝑘𝑘𝐵𝐵 𝑇𝑇 = 𝑁𝑁𝑘𝑘𝐵𝐵 𝑇𝑇 = 𝑛𝑛𝑛𝑛𝑛𝑛
2 2 2
1 𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 5 7
𝐶𝐶𝑉𝑉 = = 𝑅𝑅, 𝐶𝐶𝑃𝑃 = 𝐶𝐶𝑉𝑉 + 𝑅𝑅 = 𝑅𝑅
𝑛𝑛 𝑑𝑑𝑑𝑑 2 2
𝐶𝐶𝑃𝑃 7
𝛾𝛾𝑑𝑑𝑑𝑑𝑑𝑑 = = = 1.4 (lower temperatures)
𝐶𝐶𝑉𝑉 5
Review
• Diatomic Molecule: At higher temperatures, two
vibrational degrees of freedom are used
3+2+2 7 7
𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑁𝑁 𝑘𝑘𝐵𝐵 𝑇𝑇 = 𝑁𝑁𝑘𝑘𝐵𝐵 𝑇𝑇 = 𝑛𝑛𝑛𝑛𝑛𝑛
2 2 2
1 𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 7 9
𝐶𝐶𝑉𝑉 = = 𝑅𝑅, 𝐶𝐶𝑃𝑃 = 𝐶𝐶𝑉𝑉 + 𝑅𝑅 = 𝑅𝑅
𝑛𝑛 𝑑𝑑𝑑𝑑 2 2
𝐶𝐶𝑃𝑃 9
𝛾𝛾𝑑𝑑𝑑𝑑𝑑𝑑 = = = 1.286 (higher temperatures)
𝐶𝐶𝑉𝑉 7
𝐷𝐷
• General molecule: 𝐶𝐶𝑉𝑉 = 𝑅𝑅, 𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉 = 𝑅𝑅
2
𝛾𝛾 = 𝐶𝐶𝑃𝑃 /𝐶𝐶𝑉𝑉 = 𝐷𝐷 + 2 /𝐷𝐷, 𝐷𝐷 =no. of degrees of freedom
• polyatomic molecule:
𝐷𝐷 = 3 + 3 + 2 (𝑛𝑛𝑛𝑛. 𝑜𝑜𝑜𝑜 𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝𝑝 𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣𝑣)
Review
• Mean free path:
1 2 𝑣𝑣�
𝑙𝑙 = 2 , 𝑓𝑓 = 2 𝜋𝜋𝜋𝜋 𝑣𝑣̅ 𝑛𝑛𝑉𝑉 =
2 𝜋𝜋𝜋𝜋 𝑛𝑛𝑉𝑉 𝑙𝑙
1 1
• Mean free time: = 2 �𝑛𝑛
𝑓𝑓 𝜋𝜋𝜋𝜋 𝑣𝑣 𝑉𝑉
• Adiabatic Compression for an ideal gas:
𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑 = −𝑃𝑃𝑃𝑃𝑃𝑃, 𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑛𝑛𝐶𝐶𝑉𝑉 𝑑𝑑𝑑𝑑
𝑛𝑛𝐶𝐶𝑉𝑉 𝑑𝑑𝑑𝑑 = −𝑃𝑃𝑃𝑃𝑃𝑃 (adiabatic process, ideal gas)
𝑃𝑃𝑉𝑉 𝛾𝛾 = 𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐.
𝛾𝛾−1 𝛾𝛾−1
𝑇𝑇𝑖𝑖 𝑉𝑉𝑖𝑖 = 𝑇𝑇𝑓𝑓 𝑉𝑉𝑓𝑓
Review
• Boltzmann factor: prob ∝ 𝑒𝑒 −𝐸𝐸/𝑘𝑘𝐵𝐵 𝑇𝑇
• 𝑛𝑛𝐸𝐸 (𝐸𝐸) = number of molecules per unit
volume, per unit energy interval, interval
between 𝐸𝐸 to 𝐸𝐸 + 𝑑𝑑𝑑𝑑.
• 𝑛𝑛𝐸𝐸 𝐸𝐸 = 𝑛𝑛0 𝑒𝑒 −𝐸𝐸/𝑘𝑘𝐵𝐵 𝑇𝑇
• 𝑛𝑛𝐸𝐸 𝐸𝐸 𝑑𝑑𝑑𝑑/𝑁𝑁 = 𝑁𝑁𝑣𝑣 𝑣𝑣 𝑑𝑑𝑑𝑑/𝑁𝑁 = probability of
having a molecule speed between 𝑣𝑣 and
𝑣𝑣 + 𝑑𝑑𝑑𝑑 = probability of having energy in
between 𝐸𝐸 to 𝐸𝐸 + 𝑑𝑑𝑑𝑑.
Review
• Maxwell–Boltzmann speed distribution
function:
3
𝑚𝑚 2
2 −𝑚𝑚𝑣𝑣 2 /2𝑘𝑘𝐵𝐵 𝑇𝑇
𝑁𝑁𝑣𝑣 𝑣𝑣 = 4𝜋𝜋𝜋𝜋 𝑣𝑣 𝑒𝑒
2 𝜋𝜋𝑘𝑘𝐵𝐵 𝑇𝑇
• 𝑓𝑓 𝑣𝑣 = 𝑁𝑁𝑣𝑣 𝑣𝑣 /𝑁𝑁, probability density =
probability per unit interval of speed

∞ ∞ ∫0 𝑁𝑁𝑣𝑣 𝑣𝑣 𝑑𝑑𝑑𝑑
• ∫0 𝑓𝑓 𝑣𝑣 𝑑𝑑𝑑𝑑 = ∫0 𝑑𝑑𝑑𝑑 =
𝑁𝑁
=1
Review

• Average speed: 𝑣𝑣̅ = ∫0 𝑣𝑣 𝑁𝑁𝑣𝑣 𝑣𝑣 𝑑𝑑𝑑𝑑 /𝑁𝑁
8𝑘𝑘𝐵𝐵 𝑇𝑇
𝑣𝑣𝑎𝑎𝑎𝑎𝑎𝑎 =
𝜋𝜋𝜋𝜋
∞ ∞
• RMS speed:𝑣𝑣𝑟𝑟𝑟𝑟𝑟𝑟
2 = ∫𝑣𝑣=0 𝑓𝑓 𝑣𝑣 𝑣𝑣 2 𝑑𝑑𝑑𝑑/ ∫𝑣𝑣=0 𝑓𝑓 𝑣𝑣 𝑑𝑑𝑑𝑑
3𝑘𝑘𝐵𝐵 𝑇𝑇
𝑣𝑣𝑟𝑟𝑟𝑟𝑟𝑟 =
𝑚𝑚
𝜕𝜕 2𝑘𝑘𝐵𝐵 𝑇𝑇
• Most Probable Speed: 𝑓𝑓 𝑣𝑣 |𝑣𝑣𝑝𝑝 = 0, 𝑣𝑣𝑝𝑝 =
𝜕𝜕𝜕𝜕 𝑚𝑚
Review
• 𝑁𝑁𝐸𝐸 𝐸𝐸 𝑑𝑑𝑑𝑑 = 𝑁𝑁𝑣𝑣 𝑣𝑣 𝑑𝑑𝑑𝑑
3
𝑚𝑚 2 2 −𝑚𝑚𝑣𝑣 2 /2𝑘𝑘 𝑇𝑇
• 𝑁𝑁𝑣𝑣 𝑣𝑣 = 4𝜋𝜋𝜋𝜋 𝑣𝑣 𝑒𝑒 𝐵𝐵
2 𝜋𝜋𝑘𝑘𝐵𝐵 𝑇𝑇
1
2𝑁𝑁 1
• 𝑁𝑁𝐸𝐸 𝐸𝐸 = 3 𝐸𝐸 𝑒𝑒 −𝐸𝐸/𝑘𝑘𝐵𝐵𝑇𝑇
2
𝜋𝜋
𝑘𝑘𝐵𝐵 𝑇𝑇
2
∞ ∞ 3
• 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝐸𝐸𝑡𝑡𝑡𝑡𝑡𝑡 = ∫0 𝐸𝐸 𝑑𝑑𝑑𝑑 = ∫0 𝐸𝐸𝑁𝑁𝐸𝐸 𝐸𝐸 𝑑𝑑𝑑𝑑 = 𝑁𝑁𝑁𝑁𝐵𝐵 𝑇𝑇
2
3
• �
𝐸𝐸 = 𝐸𝐸𝑡𝑡𝑡𝑡𝑡𝑡 /𝑁𝑁 = 𝑘𝑘𝐵𝐵 𝑇𝑇
2
1
• 𝐸𝐸𝑝𝑝 = 𝑘𝑘𝐵𝐵 𝑇𝑇
2
Review
• The net work done in a cyclic process is the area
enclosed by the curve representing the process.
• Efficiency of a Heat Engine: Thermal efficiency e of a
heat engine
𝑊𝑊𝑒𝑒𝑒𝑒𝑒𝑒 𝑄𝑄ℎ − |𝑄𝑄𝑐𝑐 | |𝑄𝑄𝑐𝑐 |
𝑒𝑒 = = =1−
|𝑄𝑄ℎ | |𝑄𝑄ℎ | |𝑄𝑄ℎ |
= the ratio of what you gain (work) to what you
give (energy transfer at the higher temperature)
• Coefficient of performance (COP)=
energy transferred at high temperature |𝑄𝑄ℎ |
= =
work done by heat pump 𝑊𝑊
Review
• Efficiency of the Carnot Engine:
𝑊𝑊𝑒𝑒𝑒𝑒𝑒𝑒 What you gain 𝑄𝑄ℎ −|𝑄𝑄𝑐𝑐 | Q𝑐𝑐
𝑒𝑒 = = = =1−
|𝑄𝑄ℎ | What you give |𝑄𝑄ℎ | 𝑄𝑄ℎ
|𝑄𝑄𝑐𝑐 | 𝑇𝑇𝑐𝑐
• Carnot Efficiency: 𝑒𝑒𝐶𝐶 = 1 − =1−
|𝑄𝑄ℎ | 𝑇𝑇ℎ
• Entropy:
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑄𝑄𝑟𝑟𝑟𝑟𝑟𝑟 /𝑇𝑇
𝑆𝑆 = 𝑘𝑘𝐵𝐵 ln 𝑤𝑤
𝑤𝑤=multiplicity of arrangements
Review
• For a reversible process between initial state 𝑖𝑖 and final
state 𝑓𝑓:
𝑓𝑓 𝑓𝑓 𝑑𝑑𝑄𝑄𝑟𝑟𝑟𝑟𝑟𝑟
Δ𝑆𝑆 = 𝑆𝑆𝑓𝑓 − 𝑆𝑆𝑖𝑖 = ∫𝑖𝑖 𝑑𝑑𝑑𝑑 = ∫𝑖𝑖
𝑇𝑇
• Entropy change in a reversible cycle:
� 𝑑𝑑𝑄𝑄𝑟𝑟𝑟𝑟𝑟𝑟 /𝑇𝑇 = � 𝑑𝑑𝑑𝑑 = 0

• Entropy Change for an Ideal Gas:


𝑉𝑉𝑓𝑓 𝑇𝑇𝑓𝑓
• Δ𝑆𝑆 = 𝑆𝑆𝑓𝑓 − 𝑆𝑆𝑖𝑖 = 𝑛𝑛𝑛𝑛 ln + 𝑛𝑛𝐶𝐶𝑉𝑉 ln
𝑉𝑉𝑖𝑖 𝑇𝑇𝑖𝑖
𝑁𝑁!
• 𝑤𝑤 = (multiplicity of configuration)
𝑛𝑛1 !𝑛𝑛2 !
Problems
• Problem-1: A sample of gas undergoes a transition
from an initial state a to a final state b by three
different paths (processes), as shown in the PV
diagram, where 𝑉𝑉𝑏𝑏 = 5.00𝑉𝑉𝑖𝑖 . The energy transferred to
the gas as heat in process 1 is
10𝑝𝑝𝑖𝑖 𝑉𝑉𝑖𝑖 . In terms of 𝑝𝑝𝑖𝑖 𝑉𝑉𝑖𝑖 ,what are
(a) the energy transferred
to the gas as heat in process 2
and
(b) the change in internal energy
that the gas undergoes in process 3?
Problems
• Problem-2: Show that the internal energy of a
material, whose equation of state can be
written as 𝑃𝑃 = 𝑓𝑓 𝑉𝑉 𝑇𝑇, is independent of the
volume. Here, 𝑃𝑃 is the pressure, 𝑇𝑇 is the
temperature in absolute temperature scale
and 𝑓𝑓(𝑉𝑉) is a function of temperature only.
• Solution-2:
Problems
• Solution-2: From first law, 𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
• But for a reversible process, 𝑑𝑑𝑑𝑑 = 𝑇𝑇 𝑑𝑑𝑑𝑑
• Hence, 𝑇𝑇𝑇𝑇𝑇𝑇 = 𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 − 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 + 𝑃𝑃𝑃𝑃𝑃𝑃
1
• 𝑑𝑑𝑑𝑑 = (𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 + 𝑃𝑃𝑃𝑃𝑃𝑃) ,
𝑇𝑇
𝜕𝜕𝑆𝑆 1 𝜕𝜕𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 𝑃𝑃
• ⇒ = + , again
𝜕𝜕𝜕𝜕 𝑇𝑇 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝑇𝑇
𝜕𝜕𝑆𝑆 1 𝜕𝜕𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖
• =
𝜕𝜕𝑇𝑇 𝑉𝑉 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉
• But 𝑆𝑆 is a smooth function of 𝑉𝑉, 𝑇𝑇. Hence the order of partial
derivatives in a second order mixed derivative does not matter.
𝜕𝜕2 𝑆𝑆 𝜕𝜕2 𝑆𝑆
• Hence, = , condition that 𝑑𝑑𝑑𝑑 is an exact differential.
𝜕𝜕𝑉𝑉𝑉𝑉𝑉𝑉 𝜕𝜕𝑇𝑇𝜕𝜕𝑉𝑉
𝜕𝜕 1 𝜕𝜕𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 𝜕𝜕 𝑃𝑃 𝜕𝜕 1 𝜕𝜕𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖
• + =
𝜕𝜕𝑇𝑇 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑇𝑇 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝑇𝑇 𝑇𝑇
Problems
• Solution-2: Again, 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 is a smooth function of the variables 𝑃𝑃, 𝑉𝑉,
and 𝑇𝑇.
𝜕𝜕2 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 𝜕𝜕2 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖
• Hence, =
𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕
−1 𝜕𝜕𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 1 𝜕𝜕2 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 𝜕𝜕 𝑃𝑃
• Hence, 2 + +
𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝑇𝑇 𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑇𝑇 𝑉𝑉
𝜕𝜕 1 𝜕𝜕𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 1 𝜕𝜕2 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖
= =
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝑇𝑇 𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕
𝜕𝜕𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 2 𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕
• ⇒ = 𝑇𝑇 = 𝑇𝑇 − 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕
• Since 𝑃𝑃 = 𝑓𝑓 𝑉𝑉 𝑇𝑇, = 𝑓𝑓 𝑉𝑉 = 𝑃𝑃/𝑇𝑇
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖
• Hence, = 0, i.e. 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 does not depend on 𝑉𝑉. Proved.
𝜕𝜕𝜕𝜕 𝑇𝑇
Problems
• Problem-3: Show, using the second law of
thermodynamics, that any flow of heat between
two heat reservoirs between temperatures 𝑇𝑇𝐻𝐻
and 𝑇𝑇𝐶𝐶 , where 𝑇𝑇𝐻𝐻 > 𝑇𝑇𝐶𝐶 must be from the hotter
to the cooler reservoir.
• Solution-3: A heat reservoir, is by definition a
body with infinite heat capacity.
• When a finite quantity of heat is added to
removed from the reservoir, it undergoes a finite
entropy change at constant temperature.
Problems
𝑄𝑄
• Solution-3: Δ𝑆𝑆 =
𝑇𝑇
• Heat exchanged has the same magnitude but opposite sign.
𝑄𝑄𝐻𝐻 = −𝑄𝑄𝑐𝑐
𝑄𝑄𝐻𝐻 −𝑄𝑄𝐶𝐶
• Hence, entropy changes are: Δ𝑆𝑆𝐻𝐻 = =
𝑇𝑇𝐻𝐻 𝑇𝑇𝐻𝐻
𝑄𝑄𝐶𝐶
• Δ𝑆𝑆𝐶𝐶 = ,
𝑇𝑇𝐶𝐶
1 1 𝑇𝑇𝐻𝐻 −𝑇𝑇𝐶𝐶
• Δ𝑆𝑆𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡 = Δ𝑆𝑆𝐻𝐻 +Δ𝑆𝑆𝐶𝐶 = 𝑄𝑄𝐶𝐶 − = 𝑄𝑄𝐶𝐶
𝑇𝑇𝑐𝑐 𝑇𝑇𝐻𝐻 𝑇𝑇𝐻𝐻 𝑇𝑇𝐶𝐶
• According to the second law, Δ𝑆𝑆𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡𝑡 ≥ 0
• Since, 𝑇𝑇𝐻𝐻 ≠ 𝑇𝑇𝐶𝐶 , we get 𝑄𝑄𝐶𝐶 𝑇𝑇𝐻𝐻 − 𝑇𝑇𝐶𝐶 > 0
• Hence, 𝑄𝑄𝐶𝐶 must be positive and must represent heat added to the
cold reservoir at 𝑇𝑇𝐶𝐶 .
• Thus heat must flow from the higher temperature reservoir to the
lower temperature reservoir.
Problems
• Problem 4:Two kinds of ideal gases at equal pressure
and temperature, initially separated in two containers,
are mixed by diffusion. Show that the entropy is
increased in this process by an amount
𝑛𝑛1 𝑛𝑛2
Δ𝑆𝑆 = −𝑅𝑅 𝑛𝑛1 log + 𝑛𝑛2 log
(𝑛𝑛1 +𝑛𝑛2 ) 𝑛𝑛1 + 𝑛𝑛2
where 𝑛𝑛1 and 𝑛𝑛2 are the moles of component gases.
Assume that no change in pressure and temperature
occurs due to the diffusion and that the partial pressure
of each gas in the mixture is proportional to the molar
concentration.
Problems
• Solution-10: Initially the gases 1 and 2 had the
volumes 𝑉𝑉1 = 𝑛𝑛1 𝑅𝑅𝑅𝑅/𝑝𝑝 and 𝑉𝑉2 = 𝑛𝑛2 𝑅𝑅𝑅𝑅/𝑝𝑝 and
were separated by a wall.
• When the wall is removed, they mix by diffusion.
• Instead of considering the. diffusion process
itself, we consider first a process which reversibly
separates the mixed gas in the state (𝑉𝑉 = 𝑉𝑉1 +
𝑉𝑉2 , 𝑇𝑇, 𝑝𝑝) into pure gases, one in the state
(𝑉𝑉, 𝑇𝑇, 𝑝𝑝1 ) and the other in the state (𝑉𝑉, 𝑇𝑇, 𝑝𝑝2 ) .
Problems
• Solution-4:
• Two containers of volume V, are
superimposed where the wall b is
semipermeable allowing only the gas 2 to pass
while the wall a is semi-permeable to the gas
1.
• The pressure of the impermeable gas on the
semi-permeable wall is equal to the partial
pressure of that gas.
Problems
• Solution-4: Hence, as the two containers are
drawn apart, the walls of the container with the
gas 1 experience the partial pressure 𝑝𝑝1
• This gives rise to no resultant force on the
container, and so no work is done adiabatically,
the internal energy and the temperature remains
unchanged.
• In order to bring the gases back to the initial
states, the separate gases are isothermally
compressed from volume V to 𝑉𝑉1 and 𝑉𝑉2
respectively.
Problems
• Solution-4:
𝑑𝑑𝑑𝑑 𝑝𝑝𝑝𝑝𝑝𝑝
𝑛𝑛𝑛𝑛𝑛𝑛𝑛𝑛
• By integrating, 𝑑𝑑𝑑𝑑 = = = we get,
𝑇𝑇 𝑇𝑇 𝑉𝑉
𝑉𝑉1 𝑉𝑉2
• Δ𝑆𝑆1 = 𝑛𝑛1 𝑅𝑅 log , Δ𝑆𝑆2 = 𝑛𝑛2 𝑅𝑅 log
𝑉𝑉 𝑉𝑉
• Δ𝑆𝑆1 + Δ𝑆𝑆2 is the change in entropy for reversibly
separating the mixed gases.
• Δ𝑆𝑆 = −Δ𝑆𝑆1 − Δ𝑆𝑆2
𝑛𝑛1 𝑛𝑛2
• = −𝑅𝑅 𝑛𝑛1 log + 𝑛𝑛2 log
(𝑛𝑛1 +𝑛𝑛2 ) 𝑛𝑛1 +𝑛𝑛2
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Simple Thermodynamic Systems
• Thermodynamic Variables: The coordinates necessary
and sufficient for a macroscopic description of a system
are its thermodynamic variables.
• Examples:
• A. For an ideal gas in piston-cylinder system these are
pressure, volume and temperature i.e. 𝑃𝑃, 𝑉𝑉, 𝑇𝑇.
• B. For a stretched wire, these are:
1. The tension in the wire 𝑇𝑇, measured in Newtons (N).
2. The length of the wire 𝐿𝐿, measured in meters (m).
3. The absolute temperature 𝑇𝑇, measured in kelvin (K).
Simple Thermodynamic Systems
• C. For a thin surface of liquid (e.g. a thin film of oil on the
surface of water, a soap bubble, etc.), the variables are:
1. The surface tension 𝛾𝛾 , measured in newtons per meter
(N/m).
2. The area of the film 𝐴𝐴, measured in square meters (𝑚𝑚2 ).
3. The absolute temperature 𝑇𝑇, measured in kelvin (K).
• D. For a dielectric slab inside a capacitor, these are:
1. The electric field 𝐸𝐸, measured in volts per meter (V /m).
2. The total polarization Ω = 𝑃𝑃 𝑉𝑉, measured in coulomb-
meters (C · m).
3. The absolute temperature 𝑇𝑇, measured in kelvin (K).
Simple Thermodynamic Systems
• E. For a paramagnetic rod inside a magnetic field,
these are:
1. The magnetic field 𝐻𝐻, measured in amperes
per meter (A/m).
2. The total magnetization Χ = 𝑀𝑀𝑉𝑉, measured in
ampere-meter squared (𝐴𝐴. 𝑚𝑚2 ).
3. The absolute temperature 𝑇𝑇, measured in
kelvin (K).
Simple Thermodynamic Systems
• Change of State: When the macroscopic
coordinates of a system change in any way
whatsoever, either spontaneously or by virtue of
outside influence, the system is said to undergo a
change of state.
• Isolated System: When a system is not influenced
in any way by its surroundings, it is said to be
isolated.
• In practical applications of thermodynamics,
isolated systems are of little importance.
Simple Thermodynamic Systems
• We usually have to deal with a system that is
influenced in some way by its surroundings.
• In general the surroundings may exert forces
on the system or provide contact between the
system and a body at some definite
temperature.
• When the state of a system changes,
interactions usually take place between the
system and its surroundings.
Thermodynamic Equilibrium
• Mechanical Equilibrium: When there is no
unbalanced force or torque in the interior of a
system and also none between a system and its
surroundings, the system is said to be in a state of
mechanical equilibrium.
• When these conditions are not satisfied, either
the system alone or both the system and its
surroundings will undergo a change of state.
• The change of state will cease only when
mechanical equilibrium is restored.
Thermodynamic Equilibrium
• Chemical Equilibrium: When a system in
mechanical equilibrium does not tend to
undergo:
• A. A spontaneous change of internal structure,
such as what happens in a chemical reaction, or
• B. A transfer of matter from one part of the
system to another, such as what happens in
diffusion or formation of solution, however
slowly,
− then it is said to be in a state of chemical
equilibrium.
Thermodynamic Equilibrium
• A system not in chemical equilibrium, must
undergo a change of state that, however
slowly or fast it may be.
• In some cases, the change of state is
exceedingly slow.
• The change ceases when chemical equilibrium
is reached.
Thermodynamic Equilibrium
• Thermal Equilibrium: Thermal equilibrium exists
when there is no spontaneous change in the
coordinates, of a system in mechanical and
chemical equilibrium, when it is separated from
its surroundings by diathermic walls.
• In other words, there is no exchange of heat
between the system and its surroundings.
• In thermal equilibrium, all parts of a system are at
the same temperature, and this temperature is
the same as that of the surroundings.
Thermodynamic Equilibrium
• When the conditions for all three types of
equilibrium are satisfied, the system is said to be
in a state of thermodynamic equilibrium.
• Thermodynamic Equilibrium ≡
Mechanical Equilibrium + Chemical Equilibrium
+ Thermal Equilibrium
• There will be no tendency, whatsoever, for any
change of state, either of the system or of the
surroundings, to occur when the system is in
thermodynamic equilibrium with the
surroundings.
Thermodynamic Equilibrium
• States of thermodynamic equilibrium can be
described in terms of macroscopic
coordinates that do not involve time.
• Thermodynamic coordinates do not involve
time.
• Thermodynamics does not attempt to deal
with any problem involving the rate at which a
process takes place.
Thermodynamic Equilibrium
• When the conditions for anyone of the three
types of equilibrium, that constitute
thermodynamic equilibrium are not satisfied, the
system is said to be in a non-equilibrium state.
• When the conditions for mechanical and thermal
equilibrium are not satisfied, the states traversed
by a system cannot be described in terms of
thermodynamic coordinates referring to the
system as a whole.
Equation of State
• Consider a gas of constant mass that is in a closed
system.
• The three thermodynamic coordinates, 𝑃𝑃, 𝑉𝑉 and
𝑇𝑇 are not all independent.
• Only two of these may be changed
independently.
• This implies that there exists an equation, of
equilibrium state, which connects the
thermodynamic coordinates and makes one of
these dependent on the two others.
• Such an equation, called an equation of state.
Equation Of State
• An equation of state is a mathematical function
relating the appropriate thermodynamic
coordinates of a system in equilibrium.
• Every thermodynamic system has its own
equation of state.
• In some cases the equation of state may be so
complicated that it cannot be expressed in terms
of simple mathematical functions.
• For a closed system, the equation of state relates
the temperature to two other thermodynamic
variables.
Equation Of State
• Example-1: A system consisting of a gas at
very low pressure has the simple equation of
state of an ideal gas:
𝑃𝑃𝑃𝑃 = 𝑛𝑛𝑛𝑛𝑛𝑛
• 𝑛𝑛 = the number of moles, 𝑅𝑅 = molar gas
constant.
𝑃𝑃𝑣𝑣 = 𝑅𝑅𝑅𝑅
• 𝑣𝑣 = molar volume= 𝑉𝑉/𝑛𝑛
Equation Of State
• Example-2: At higher pressures, the equation
of state is more complicated, being fairly well
represented by the van der Waals equation,
which takes into account particle interactions
and the finite size of the particles:
𝑎𝑎
𝑃𝑃 + 2 𝑣𝑣 − 𝑏𝑏 = 𝑅𝑅𝑅𝑅
𝑣𝑣
where 𝑎𝑎 and 𝑏𝑏 are gas constants appropriate
for any particular gas.
Hydrostatic Systems
• Definition: Any isotropic system of constant mass
and constant composition that exerts on the
surroundings a uniform hydrostatic pressure, in
the absence of gravitational, electric, and
magnetic effects is called a hydrostatic system.
• Categories of Hydrostatic Systems:
1. Pure System
2. Homogeneous Mixture.
3. Heterogeneous Mixture.
Hydrostatic Systems
• Pure System: A pure substance, which is a single
chemical compound in the form of a solid, a
liquid, a gas, a mixture of any two, or a mixture of
all three (solid, liquid or gas).
• Homogeneous Mixture: A homogeneous mixture
of different compounds, such as a mixture of inert
gases, a mixture of chemically active gases, a
mixture of liquids, or a solution.
• Heterogeneous Mixture: A heterogeneous
mixture, such as a mixture of different gases in
contact with a mixture of different liquids.
Hydrostatic Systems
• For a system in single phase, a hydrostatic
system can be described by an equation of
state having three coordinates, pressure 𝑃𝑃,
volume 𝑉𝑉 and temperature 𝑇𝑇 (an
experimental fact).
• Every hydrostatic system, that is, a 𝑃𝑃𝑃𝑃𝑃𝑃
system, has an equation of state expressing a
relation among these three coordinates that is
valid for all equilibrium states.
Infinitesimal Quantities
• Infinitesimal volume 𝑑𝑑𝑉𝑉 : An infinitesimal volume 𝑑𝑑𝑉𝑉
must be:
small compared to the macroscopic volume
large compared to the volume of the particles
• 𝑑𝑑𝑉𝑉 must contain a large number of particles.
• Infinitesimal pressure 𝑑𝑑𝑃𝑃: An infinitesimal pressure 𝑑𝑑𝑃𝑃
must be:
small compared to the total pressure of the system
large compared to the local fluctuations of pressure caused
by the momentary variations of the microscopic concentrations.
Infinitesimal Quantities
• Every infinitesimal in thermodynamics must
satisfy the requirement that:
• A. it represents a change in a quantity which is
small with respect to the quantity itself and
• B. large in comparison with the effect produced
by the behavior of a few molecules.
• The reason for this is that thermodynamic
coordinates such as volume, pressure, and
temperature have no meaning when applied to a
few particles (i.e. these are macroscopic
coordinates).
Equation of State of a Hydrostatic
System
• Equation of state of a hydrostatic system may
be written as:
𝑉𝑉 = 𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹 𝑇𝑇, 𝑃𝑃 = 𝑉𝑉(𝑇𝑇, 𝑃𝑃)
Or,
𝑃𝑃 = 𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹 𝑇𝑇, 𝑉𝑉 = 𝑃𝑃(𝑇𝑇, 𝑉𝑉)
Or,
𝑇𝑇 = 𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹𝐹 𝑃𝑃, 𝑉𝑉 = 𝑇𝑇(𝑃𝑃, 𝑉𝑉)
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Hydrostatic Systems
• Definition: Any isotropic system of constant mass
and constant composition that exerts on the
surroundings a uniform hydrostatic pressure, in
the absence of gravitational, electric, and
magnetic effects is called a hydrostatic system.
• A hydrostatic system may be described by three
macroscopic quantities 𝑃𝑃, 𝑉𝑉 and 𝑇𝑇.
• Hence a hydrostatic system is often called a 𝑃𝑃𝑃𝑃𝑃𝑃
system.
Hydrostatic Systems
• Pure System: A pure substance, which is a single
chemical compound in the form of a solid, a
liquid, a gas, a mixture of any two, or a mixture of
all three (solid, liquid or gas).
• Homogeneous Mixture: A homogeneous mixture
of different compounds, such as a mixture of inert
gases, a mixture of chemically active gases, a
mixture of liquids, or a solution.
• Heterogeneous Mixture: A heterogeneous
mixture, such as a mixture of different gases in
contact with a mixture of different liquids.
Infinitesimal Quantities
• Every infinitesimal in thermodynamics must
satisfy the requirement that:
• A. it represents a change in a quantity which is
small with respect to the quantity itself and
• B. large in comparison with the effect produced
by the behavior of a few molecules.
• The reason for this is that thermodynamic
coordinates such as volume, pressure, and
temperature have no meaning when applied to a
few particles (i.e. these are macroscopic
coordinates).
Infinitesimals of Hydrostatic Systems
• We may write the equation of a hydrostatic
system as:
• 𝑉𝑉 = 𝑉𝑉(𝑇𝑇, 𝑃𝑃) or, 𝑃𝑃 = 𝑃𝑃(𝑉𝑉, 𝑇𝑇) or, 𝑇𝑇 = 𝑇𝑇(𝑃𝑃, 𝑉𝑉)
• Hence we may write:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
• A. 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑 , where,
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕
• (i) the partial derivative implies an infinitesimal
𝜕𝜕𝜕𝜕 𝑃𝑃
change in 𝑉𝑉 for an infinitesimal change in 𝑇𝑇 with 𝑃𝑃 held
constant during the operation of differentiation and
𝜕𝜕𝜕𝜕
likewise for .
𝜕𝜕𝜕𝜕 𝑇𝑇
Infinitesimals of Hydrostatic Systems
𝜕𝜕𝜕𝜕
• (ii) Note that, both partial derivatives, and
𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕
can be functions of 𝑇𝑇 and 𝑃𝑃.
𝜕𝜕𝜕𝜕 𝑇𝑇
• (iii) Recall, the quantity called the average coefficient
of volume expansion:
Average coefficient of volume expansion
Change of volume per unit volume
=
Change of temperature
where the change occurs at constant pressure.
Infinitesimals of Hydrostatic Systems
• In the infinitesimal limit, the average coefficient
of volume expansion becomes what is called the
volume expansivity:
1 𝜕𝜕𝜕𝜕
𝛽𝛽 =
𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃
• 𝛽𝛽 is always positive except for a few notable
exceptions: liquid water between 0°C and 4°C
experiences a decrease in V when T increases,
and so too for an ordinary rubber band at room
temperature.
Infinitesimals of Hydrostatic Systems
• (iv) Recall the quantity called the average bulk modulus:
Average bulk modulus
Change of pressure
=−
Change of volume per unit volume
• If we further require that the temperature be kept constant, then
we define isothermal bulk modulus as:
𝜕𝜕𝜕𝜕
𝐵𝐵 = −𝑉𝑉
𝜕𝜕𝜕𝜕 𝑇𝑇
• The inverse of 𝐵𝐵 is defined as the isothermal compressibility as:
1 𝜕𝜕𝑉𝑉
𝜅𝜅 = −
𝑉𝑉 𝜕𝜕𝑃𝑃 𝑇𝑇
Infinitesimals of Hydrostatic Systems
• Again we may write the equation of a hydrostatic system
as:
𝑃𝑃 = 𝑃𝑃(𝑇𝑇, 𝑉𝑉)
𝜕𝜕𝑃𝑃 𝜕𝜕𝑃𝑃
• B. Hence 𝑑𝑑𝑃𝑃 = 𝜕𝜕𝜕𝜕 𝑑𝑑𝑑𝑑 + 𝜕𝜕𝑉𝑉 𝑑𝑑𝑉𝑉 , where,
𝑉𝑉 𝑇𝑇
𝜕𝜕𝑃𝑃
• (i) the partial derivative implies an infinitesimal
𝜕𝜕𝜕𝜕 𝑉𝑉
change in 𝑃𝑃 for an infinitesimal change in 𝑇𝑇 with 𝑉𝑉 held
constant during the operation of differentiation, and
𝜕𝜕𝜕𝜕
likewise for .
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
• Note that, both partial derivatives, and can be
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑇𝑇
functions of 𝑇𝑇 and 𝑉𝑉.
Infinitesimals of Hydrostatic Systems
• Again we may write the equation of a hydrostatic system
as:
𝑇𝑇 = 𝑇𝑇(𝑃𝑃, 𝑉𝑉)
𝜕𝜕𝑇𝑇 𝜕𝜕𝑇𝑇
• C. Hence 𝑑𝑑𝑇𝑇 = 𝜕𝜕𝑃𝑃 𝑑𝑑𝑃𝑃 + 𝜕𝜕𝑉𝑉 𝑑𝑑𝑉𝑉 , where,
𝑉𝑉 𝑃𝑃
𝜕𝜕𝑇𝑇
• (i) the partial derivative implies an infinitesimal
𝜕𝜕𝑃𝑃 𝑉𝑉
change in 𝑇𝑇 for an infinitesimal change in 𝑃𝑃 with 𝑉𝑉 held
constant during the operation of differentiation, and
𝜕𝜕𝑇𝑇
likewise for .
𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
• Note that, both partial derivatives, and can be
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃
functions of 𝑃𝑃 and 𝑉𝑉.
Exact and Inexact Differentials
• Exact Differentials: The differentials 𝑑𝑑𝑑𝑑, 𝑑𝑑𝑑𝑑, and 𝑑𝑑𝑑𝑑 are
differentials of actual functions 𝑃𝑃 = 𝑃𝑃(𝑉𝑉, 𝑇𝑇), V = V(P, T)
and 𝑇𝑇 = 𝑇𝑇(𝑃𝑃, 𝑉𝑉) and are called exact differentials.
• An exact differential may be expressed as:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑦𝑦
𝜕𝜕𝜕𝜕 𝑦𝑦 𝜕𝜕𝑦𝑦 𝑥𝑥
where, 𝑧𝑧 = 𝑧𝑧(𝑥𝑥, 𝑦𝑦).
• Inexact Differentials: An infinitesimal that is not an
differential of an actual function, is called an inexact
differential and cannot be expressed by an equation of the
above type.
• Example: Infinitesimal heat transferred 𝑑𝑑𝑑𝑑 or work done 𝑑𝑑𝑑𝑑.
Mathematical Theorems
• Theorem-1: Suppose that there exists a functional
relationship among the three coordinates 𝑥𝑥, 𝑦𝑦, and 𝑧𝑧, i.e.
𝑓𝑓 𝑥𝑥, 𝑦𝑦, 𝑧𝑧 = 0
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
then =1.
𝜕𝜕𝜕𝜕 𝑧𝑧 𝜕𝜕𝜕𝜕 𝑧𝑧
• Proof: There exists a functional relationship among the
three coordinates 𝑥𝑥, 𝑦𝑦, and 𝑧𝑧, i.e.
𝑓𝑓 𝑥𝑥, 𝑦𝑦, 𝑧𝑧 = 0
• Then, 𝑥𝑥 can be imagined as a function of 𝑦𝑦 and 𝑧𝑧, i.e.
𝑥𝑥 = 𝑥𝑥(𝑦𝑦, 𝑧𝑧). Hence,
𝜕𝜕𝑥𝑥 𝜕𝜕𝑥𝑥
𝑑𝑑𝑥𝑥 = 𝑑𝑑𝑦𝑦 + 𝑑𝑑𝑧𝑧
𝜕𝜕𝑦𝑦 𝑧𝑧 𝜕𝜕𝑧𝑧 𝑦𝑦
Mathematical Theorems
• Again, 𝑦𝑦 can be imagined as a function of 𝑥𝑥 and 𝑧𝑧,
i.e. 𝑦𝑦 = 𝑦𝑦(𝑥𝑥, 𝑧𝑧). Hence,
𝜕𝜕𝑦𝑦 𝜕𝜕𝑦𝑦
𝑑𝑑𝑦𝑦 = 𝑑𝑑𝑥𝑥 + 𝑑𝑑𝑧𝑧
𝜕𝜕𝑥𝑥 𝑧𝑧 𝜕𝜕𝑧𝑧 𝑥𝑥
• Hence, substituting the second into the first we
get:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑧𝑧 𝜕𝜕𝜕𝜕 𝑧𝑧 𝜕𝜕𝜕𝜕 𝑥𝑥 𝜕𝜕𝜕𝜕 𝑦𝑦
Mathematical Theorems
• Hence,
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑧𝑧
𝜕𝜕𝜕𝜕 𝑧𝑧
𝜕𝜕𝜕𝜕 𝑧𝑧
𝜕𝜕𝜕𝜕 𝑥𝑥
𝜕𝜕𝜕𝜕
+ 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑦𝑦
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
= 𝑑𝑑𝑑𝑑 + + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑧𝑧
𝜕𝜕𝜕𝜕 𝑧𝑧
𝜕𝜕𝜕𝜕 𝑧𝑧
𝜕𝜕𝜕𝜕 𝑥𝑥
𝜕𝜕𝜕𝜕 𝑦𝑦
• Of the three coordinates, only two are independent.
Mathematical Theorems
• Choosing 𝑥𝑥 and 𝑧𝑧 as the independent
coordinates, the above equation must be true
for all sets of values of 𝑑𝑑𝑑𝑑 and 𝑑𝑑𝑑𝑑.
• Thus, if 𝑑𝑑𝑑𝑑 = 0 and 𝑑𝑑𝑑𝑑 ≠ 0, it follows that
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
=1
𝜕𝜕𝜕𝜕 𝑧𝑧 𝜕𝜕𝜕𝜕 𝑧𝑧
𝜕𝜕𝜕𝜕 1
• Or, = 𝜕𝜕𝜕𝜕 Proved.
𝜕𝜕𝜕𝜕 𝑧𝑧
𝜕𝜕𝜕𝜕 𝑧𝑧
Mathematical Theorems
• Theorem-2: Suppose that there exists a functional
relationship among the three coordinates 𝑥𝑥, 𝑦𝑦, and
𝑧𝑧, i.e. 𝑓𝑓 𝑥𝑥, 𝑦𝑦, 𝑧𝑧 = 0
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
then =−
𝜕𝜕𝜕𝜕 𝑧𝑧 𝜕𝜕𝜕𝜕 𝑥𝑥 𝜕𝜕𝜕𝜕 𝑦𝑦

• Proof: Since, there exists a functional relationship


between the variables 𝑥𝑥, 𝑦𝑦 and 𝑧𝑧, we can consider 𝑥𝑥
as a function of 𝑦𝑦 and 𝑧𝑧, i.e. 𝑥𝑥 = 𝑥𝑥(𝑦𝑦, 𝑧𝑧).
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
• Hence, 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑧𝑧 𝜕𝜕𝜕𝜕 𝑦𝑦
Mathematical Theorems
• Again, 𝑦𝑦 can be imagined as a function of 𝑥𝑥 and 𝑧𝑧,
i.e. 𝑦𝑦 = 𝑦𝑦(𝑥𝑥, 𝑧𝑧). Hence,
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑧𝑧 𝜕𝜕𝜕𝜕 𝑥𝑥
• Hence, substituting the second into the first we
get:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑧𝑧 𝜕𝜕𝜕𝜕 𝑧𝑧 𝜕𝜕𝜕𝜕 𝑥𝑥 𝜕𝜕𝜕𝜕 𝑦𝑦
Mathematical Theorems
• Hence,
𝑑𝑑𝑑𝑑 =
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 + + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑧𝑧 𝜕𝜕𝜕𝜕 𝑧𝑧 𝜕𝜕𝜕𝜕 𝑧𝑧 𝜕𝜕𝜕𝜕 𝑥𝑥 𝜕𝜕𝜕𝜕 𝑦𝑦

• Of the three coordinates, only two are


independent.
• We can choose Choosing 𝑥𝑥 and 𝑧𝑧 as the
independent coordinates, the above equation
must be true for all sets of values of 𝑑𝑑𝑑𝑑 and 𝑑𝑑𝑑𝑑.
Mathematical Theorems
• Thus, if 𝑑𝑑𝑑𝑑 ≠ 0 and 𝑑𝑑𝑑𝑑 = 0, it follows that
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
+ =0
𝜕𝜕𝜕𝜕 𝑧𝑧 𝜕𝜕𝜕𝜕 𝑥𝑥 𝜕𝜕𝜕𝜕 𝑦𝑦
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
• Hence, =− Proved.
𝜕𝜕𝜕𝜕 𝑧𝑧 𝜕𝜕𝜕𝜕 𝑥𝑥 𝜕𝜕𝜕𝜕 𝑦𝑦
• Corollary-1: Combining the results of theorems 1
and 2, we get:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
= −1
𝜕𝜕𝜕𝜕 𝑧𝑧 𝜕𝜕𝜕𝜕 𝑥𝑥 𝜕𝜕𝜕𝜕 𝑦𝑦
Mathematical Theorems
• Example-1: In the case of a hydrostatic system,
i.e a 𝑃𝑃𝑃𝑃𝑃𝑃 system, we get (𝑥𝑥 = 𝑃𝑃, 𝑦𝑦 = 𝑉𝑉, 𝑧𝑧 =
𝑇𝑇)
𝜕𝜕𝑃𝑃 𝜕𝜕𝑉𝑉 𝜕𝜕𝑃𝑃
=−
𝜕𝜕𝑉𝑉 𝑇𝑇 𝜕𝜕𝑇𝑇 𝑃𝑃 𝜕𝜕𝑇𝑇 𝑉𝑉
• The volume expansivity and the isothermal
compressibility are defined as:
1 𝜕𝜕𝜕𝜕 1 𝜕𝜕𝜕𝜕
𝛽𝛽 = and 𝜅𝜅 = − .
𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑇𝑇
Mathematical Theorems
• Example-1(contd.): Hence,
1 𝜕𝜕𝜕𝜕
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 1 𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃
= −𝑉𝑉 =
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 1 𝜕𝜕𝑉𝑉
𝑉𝑉 𝑇𝑇 𝑃𝑃 −
𝑉𝑉 𝜕𝜕𝑃𝑃 𝑇𝑇
𝜕𝜕𝜕𝜕 𝛽𝛽
=
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜅𝜅
• An infinitesimal change in pressure may now be
expressed in terms of these physical quantities 𝛽𝛽 and
𝜅𝜅.
Mathematical Theorems
• Example-1(contd.): Thus,
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑇𝑇
𝛽𝛽 1
= 𝑑𝑑𝑑𝑑 − 𝑑𝑑𝑑𝑑
𝜅𝜅 𝜅𝜅𝜅𝜅
𝛽𝛽
• At constant volume, 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑
𝜅𝜅
• Hence, at constant volume, if the temperature
is changed from 𝑇𝑇𝑖𝑖 to 𝑇𝑇𝑓𝑓
Mathematical Theorems
• Example-1(contd.): Hence, at constant
volume, if the temperature is changed from 𝑇𝑇𝑖𝑖
to 𝑇𝑇𝑓𝑓 , the change in pressure would be:
𝑇𝑇𝑓𝑓
𝛽𝛽 𝛽𝛽
𝑃𝑃𝑓𝑓 − 𝑃𝑃𝑖𝑖 = � 𝑑𝑑𝑑𝑑 ≈ (𝑇𝑇𝑓𝑓 − 𝑇𝑇𝑖𝑖 )
𝑇𝑇𝑖𝑖 𝜅𝜅 𝜅𝜅
• Where we have assumed that the volume
expansivity and the isothermal compressibility
vary little if the range of temperature is small.
Mathematical Theorems
• Example-2: A mass of mercury at standard
atmospheric pressure and a temperature of 15°C
is kept at constant volume.
• If the temperature is raised to 25°C, what will be
the final pressure? (Given that 𝛽𝛽 = 1.81 ×
10−4 𝐾𝐾 −1 , 𝜅𝜅 = 4.01 × 10−11 𝑃𝑃𝑎𝑎−1 )
𝑇𝑇𝑓𝑓 𝛽𝛽 𝛽𝛽
• Solution: 𝑃𝑃𝑓𝑓 − 𝑃𝑃𝑖𝑖 = ∫𝑇𝑇 𝜅𝜅 𝑑𝑑𝑑𝑑 ≈ (𝑇𝑇𝑓𝑓 − 𝑇𝑇𝑖𝑖 )
𝑖𝑖 𝜅𝜅
1.81×10−4 𝐾𝐾−1 ×10𝐾𝐾
• Δ𝑃𝑃 = = 4.51 × 107 𝑃𝑃𝑃𝑃
4.01×10−11 𝑃𝑃𝑎𝑎−1
• With 𝑃𝑃𝑖𝑖 = 1 𝑎𝑎𝑎𝑎𝑎𝑎 = 1 × 105 𝑃𝑃𝑃𝑃, 𝑃𝑃𝑓𝑓 = 452 𝑎𝑎𝑎𝑎𝑎𝑎.
Quiz-1
• The equation of state of an ideal gas is given
as:
𝑃𝑃𝑃𝑃 = 𝑛𝑛𝑛𝑛𝑛𝑛
where 𝑛𝑛 and 𝑅𝑅 are constants.
(a) Show that the volume expansivity 𝛽𝛽 is
equal to 1/ 𝑇𝑇.
(b) Show that the isothermal compressibility 𝜅𝜅,
is equal to 1/ 𝑃𝑃.
Quiz-2
• The equation of state of a van der Waals gas is
given as:
𝑎𝑎
𝑃𝑃 + 2 𝑣𝑣 − 𝑏𝑏 = 𝑅𝑅𝑅𝑅
𝑣𝑣
where 𝑎𝑎, 𝑏𝑏, and 𝑅𝑅 are constants. Calculate the
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
following quantities: and
𝜕𝜕𝑣𝑣 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑣𝑣
𝜕𝜕𝑣𝑣
Hence calculate,
𝜕𝜕𝑇𝑇 𝑃𝑃
Quiz-2
• Solution: The van der Waals equation of state is
𝑎𝑎
𝑃𝑃 + 2 𝑣𝑣 − 𝑏𝑏 = 𝑅𝑅𝑅𝑅, which can be rewritten as:
𝑣𝑣 𝑅𝑅𝑅𝑅 𝑎𝑎
𝑃𝑃 = − 2
(𝑣𝑣−𝑏𝑏) 𝑣𝑣
𝜕𝜕𝜕𝜕 𝑅𝑅𝑅𝑅 2𝑎𝑎
• Hence, = − + 3 and
𝜕𝜕𝜕𝜕 𝑇𝑇 𝑣𝑣−𝑏𝑏 2 𝑣𝑣
𝜕𝜕𝜕𝜕 𝑅𝑅
• =
𝜕𝜕𝜕𝜕 𝑣𝑣 𝑣𝑣−𝑏𝑏
• Hence, 𝜕𝜕𝜕𝜕
𝜕𝜕𝜕𝜕 1 𝜕𝜕𝜕𝜕 𝑣𝑣 𝑅𝑅 𝑣𝑣−𝑏𝑏 𝑣𝑣 3
=− 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 =− 𝜕𝜕𝜕𝜕 =
𝜕𝜕𝜕𝜕 𝑃𝑃 𝑅𝑅𝑅𝑅𝑣𝑣 3 −2𝑎𝑎 𝑣𝑣−𝑏𝑏 2
𝜕𝜕𝜕𝜕 𝑣𝑣 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Simple Thermodynamic Systems
• We will study a number of simple
thermodynamic systems such as:
• A. Real Gas
• B. Stretched Wire
• C. Thin Surfaces
• D. Dielectric Slab in a Capacitor
• E. Paramagnetic Rod inside a Magnetic Field
Intermolecular Forces
• Intermolecular Forces: Forces between molecules play
an important role in the behavior of matter.
• Apart from gravitational attraction between the
molecules, which is very weak, intermolecular forces
are of electromagnetic in nature.
• When molecules come near each other, their electron
clouds get distorted.
• This creates induced electric dipole (and a small
amount of higher/other) moments.
• Besides, induced dipoles, certain molecules also have
permanent electric dipole moments (e.g. HCl, NH3 ,
H2 O, etc).
Intermolecular Forces
• Since molecules are in general electrically neutral
(unless there are ions), dipoles play the most
important part in creating molecular forces.
• The net force is of very short range, i.e. exist only
when they molecules are very close to each
other.
• This is a force between molecules and not
between different atoms inside a molecule
(which is intramolecular force).
Intermolecular Forces
• Attractive Molecular Force: Because of
interaction between charges inside adjacent
molecules, attractive molecular forces occur.
• This happens because, the average distance
between opposite charges between molecules
become slightly smaller than the average
distance between like (same) charges.
Intermolecular Forces
• Types of Attractive Molecular Forces: There
are four types of attractive intermolecular
forces:
• A. Dipole-Dipole Forces van der
• B. Dipole-induced Dipole Forces Waals
• C. London Dispersion Forces force
• D. Ion-Dipole Forces.
Intermolecular Forces
• Attractive Molecular Force:
• A. Dipole-Dipole Forces: Forces that act
between polar molecules.

Note: Very strong dipole-dipole bonds are called hydrogen bonds.


Intermolecular Forces
• Attractive Molecular Force:
• B. Dipole-induced Dipole Forces (Debye Forces):
This results when a polar molecule induces
a dipole in an atom or in a nonpolar molecule by
disturbing the arrangement of electrons in the
nonpolar species.
• Q. How can non-polar molecule such as O2 and
I2 dissolve in water?
• Answer: The water dipole induces a dipole in the
O2 electric cloud resulting in an attractive force.
Intermolecular Forces
• This interaction is called the Debye force,
named after Peter J.W. Debye.
• Debye Force is the angle averaged dipole-
induced dipole interaction between two
atoms or molecules and the potential varies as
− 1/𝑟𝑟 6 .
Intermolecular Forces
• Attractive Molecular Force:
• C. London Dispersion Forces: This is the attractive
force that arise as a result of temporary or
instantaneous dipoles induced in the atoms or
molecules.
• It was first described by German-American
physicist Fritz London in 1930.
• When electrons move around in an atom or
molecule, their motion is random. So there is a
chance of more electrons to be one side of a
molecule/atom than other side.
Intermolecular Forces
• The molecule/atom would develop an
instantaneous dipole moment.
Intermolecular Forces
• Strength of London Dispersion Force: It depends
on:
• A. Complexity of the Molecules:
Long chain molecules
attract much strongly
than small chain molecules.
Number of places along
the chain that can be
attracted are higher.
Intermolecular Forces
• Strength of London Dispersion Force:
• B. Molecular Size: Large molecules attract each other
more strongly than small molecules.
• As the size increases, electron cloud could be polarized
more easily and the interaction becomes stronger.
• In larger molecules, the valence electrons are farther
from the nuclei and are less tightly bound.
• Example: Noble gases are all non-polar monoatomic
molecular gases. The stronger the attractive forces
between the molecules, the higher the boiling point
will be.
Intermolecular Forces
Boiling Point of Noble Gases
Noble Gas Molar Mass (gm/mol) Boiling Point (K)
He 4.00 4.2
20.18 27.0
Ne
87.0
Ar
39.95

Kr 83.80 120.0

Xe 131.30 165.0
Intermolecular Forces
• London dispersion forces are responsible for the
fact that non-polar molecules can be condensed
to form liquids and sometimes solids.
• In polar molecules, London dispersion forces are
weaker than the dipole-dipole interaction and
hence are neglected.
• It is the weakest of all molecular forces.
• London dispersion force occurs between all
molecules, polar or non-polar.
Intermolecular Forces
• Attractive Molecular Force:
• D. Ion-Dipole Forces: These are attractive forces
between ions and polar molecules or induced
dipoles in non-polar molecules.
• These are most commonly
found in solutions. Especially
important for solutions of
ionic compounds in
polar liquids like water.
Intermolecular Forces
• A positive ion (cation) attracts the partially
negative end of a neutral polar molecule.
• A negative ion (anion) attracts the partially
positive end of a neutral polar molecule.
Van der Waals Force
• The umbrella term “van der Waals” force is used to
include:
(a) dipole-dipole force (Keesom force), ( HCl−HCl )
(b) dipole-induced dipole interaction, ( HCl−Cl2 )
(c) London dispersion force. ( Cl2 −Cl2 )
• Named after Dutch scientist Johannes Diderik van
der Waals, all three types of van der Waals forces
1
have potentials that vary as − 6 .
𝑟𝑟
Van der Waals Force
• The three types of attractive intermolecular
forces are collectively called van der Waals forces.
• The dipole-dipole interaction inversely
depends on temperature while the other
two types are independent of temperature.
• This force is the attractive part of the force
due to L-J potential.
• 𝑈𝑈vdW =
𝐶𝐶
𝑈𝑈𝐾𝐾𝐾𝐾𝐾𝐾𝐾𝐾𝐾𝐾𝐾𝐾 𝑟𝑟 + 𝑈𝑈𝐷𝐷𝐷𝐷𝐷𝐷𝐷𝐷𝐷𝐷 𝑟𝑟 + 𝑈𝑈𝐿𝐿𝐿𝐿𝐿𝐿𝐿𝐿𝐿𝐿𝐿𝐿 𝑟𝑟 = −
𝑟𝑟 6
Van der Waals Force
2 𝜇𝜇1 𝜇𝜇2 1
• 𝑈𝑈𝐾𝐾𝐾𝐾𝐾𝐾𝐾𝐾𝐾𝐾𝐾𝐾 𝑟𝑟 = −
3𝑘𝑘𝐵𝐵 𝑇𝑇 4𝜋𝜋𝜖𝜖0 𝑟𝑟 6
𝜇𝜇2 𝛼𝛼 ′ 1
• 𝑈𝑈𝐷𝐷𝐷𝐷𝐷𝐷𝐷𝐷𝐷𝐷 𝑟𝑟 = − , 𝛼𝛼 ′ = 𝛼𝛼/4 𝜋𝜋𝜖𝜖0
4𝜋𝜋𝜖𝜖0 𝑟𝑟 6
• 𝛼𝛼 ′ is polarizability volume, 𝛼𝛼 =polarizability
3 𝛼𝛼1 𝛼𝛼2 1 ℎ𝜈𝜈1 ℎ𝜈𝜈2
• 𝑈𝑈𝐿𝐿𝐿𝐿𝐿𝐿𝐿𝐿𝐿𝐿𝐿𝐿 𝑟𝑟 = − 2 6 ,
2 4𝜋𝜋𝜖𝜖0 𝑟𝑟 ℎ𝜈𝜈1 +ℎ𝜈𝜈2
𝜈𝜈𝑖𝑖 =electronic absorption frequency of i-th
molecule.
Van der Waals Force
• Geckos can stick to walls and ceilings because
of van der Waals forces.
Repulsive van der Waals force
• Recently, it has been shown that in certain situations,
van der Waals forces may become repulsive. This
happens when molecules interact in confined spaces.
• “However, in nature molecules in most cases interact in
confined spaces, such as cells, membranes, nanotubes,
etc. In is this particular situation, van der Waals forces
become repulsive at large distances between
molecules”-Prof Tkatchenko.
• Mainak Sadhukhan, Alexandre Tkatchenko. Long-Range
Repulsion Between Spatially Confined van der Waals
Dimers. Physical Review Letters, 2017; 118 (21)
DOI: 10.1103/PhysRevLett.118.210402
Intermolecular Forces
• Repulsive Molecular Force: If the molecules
come very close together, their outer charges
begin to overlap, so that the intermolecular
forces become repulsive.
• The molecules repel each other because there is
no way for a molecule to rearrange itself
internally (i.e. redistribute the electronic cloud)
to prevent repulsion of the adjacent external
electrons.
• This is pictured as repulsion on contact by billiard
ball picture of gases.
Intermolecular Forces
• Repulsive Molecular Force:
• Hence, molecules repel,
if they come too close
together (at 𝑟𝑟~1 − 2 Å),
due to electrostatic
repulsion of electron
clouds.
Intermolecular Forces
• Resultant Interatomic Force: The resultant force has a
Lennard-Jones (L-J) potential
• 𝑈𝑈 𝑟𝑟 = 𝑈𝑈𝑅𝑅𝑅𝑅𝑅𝑅𝑅𝑅𝑅𝑅 𝑟𝑟 + 𝑈𝑈𝐴𝐴𝐴𝐴𝐴𝐴𝐴𝐴𝐴𝐴𝐴𝐴𝐴𝐴 𝑟𝑟 𝐹𝐹⃗ 𝑟𝑟 = −𝑟𝑟̂ 𝑑𝑑𝑑𝑑(𝑟𝑟)/𝑑𝑑𝑑𝑑
1 𝑛𝑛 1 6
• 𝑈𝑈 𝑟𝑟 = 𝐴𝐴 − 𝐵𝐵 , 𝑛𝑛 > 9
𝑟𝑟 𝑟𝑟 Repulsive potential
𝜎𝜎 𝑛𝑛 𝜎𝜎 6
= 4𝜖𝜖 −
𝑟𝑟 𝑟𝑟
• Most commonly, 𝑛𝑛 = 12
is used.
• “6 -12” potential is used
mostly for convenience
of calculations.
Attractive potential
Real Gases
• Real gases are gases, under conditions in
which they do not obey the ideal gas law.
• It was observed experimentally that, for some
gases (e.g., CO2) at high pressures or low
temperatures, noticeable deviations from
ideal gas behavior were exhibited.
• A satisfying explanation of these deviations in
terms of intermolecular forces was first given
by van der Waals in 1873.
Real Gases
• Real gases do not obey many of the
assumptions of the kinetic theory of gases.
• A. Molecules have a finite, small but certain
nonzero volume.
• B. The range of force between the molecules
extend beyond “collision diameter” i.e forces
are not just contact force.
• In developing a modified equation of state, we
take into account the above facts.
Real Gases
• A. Finite Volume of Molecules: Consider each
molecule to be a sphere of radius 𝑟𝑟.
• Two molecules can not come closer together than
𝑑𝑑 = 2𝑟𝑟 (i.e. their separation can not be < 2𝑟𝑟).
• Around each molecule, there is an excluded
volume i.e. a region where another molecule
cannot penetrate/come.
• Excluded volume around one molecule = sphere
4
of radius 𝑑𝑑 = Ω = 𝜋𝜋𝑑𝑑 3 = 8(4𝜋𝜋𝑟𝑟 3 /3)
3
Real Gases
• A. Finite Volume of Molecules:
• But this is the excluded volume between a pair of
molecules.
• Hence, we must divide by two to avoid over-counting.
• Hence, excluded volume
′ Ω 4
for each molecule 𝑏𝑏 = = 4( 𝜋𝜋𝑟𝑟 3 )
2 3
• For one mole of molecules,
total excluded volume is
𝑏𝑏 = 𝑁𝑁𝐴𝐴 𝑏𝑏′
Note: Two molecules approaching each other the volume in
which they interact is a hemisphere not a full sphere. Hence the factor of ½
Real Gases
• A. Finite Volume of Molecules: Hence, the
excluded volume should be subtracted from the
actual volume of the container.
• The net volume is the free volume in which the
molecules could undergo translational motion.
• The net free volume = 𝑉𝑉 − 𝑛𝑛𝑛𝑛, where 𝑛𝑛 = no. of
moles of gas and 𝑏𝑏 =excluded volume per mole.
• Hence, the finite volume of gas molecule
introduces a corrected equation of state:
𝑛𝑛𝑛𝑛𝑛𝑛
𝑃𝑃 =
𝑉𝑉 − 𝑛𝑛𝑛𝑛
Real Gases
• A. Finite Volume of Molecules:
• Hence, the pressure of the real gas is greater
relative to that of an ideal gas, under the same
conditions.
• The reduced volume available to the gas
molecules implies that they make more
collisions with the walls and thereby increase
the pressure.
Real Gases
• B. Finite Range of Intermolecular Interaction:
• The attractive forces between molecules are
weaker, but have a much longer range than that
of the repulsive forces (i.e. of hard core radius).
• Consider a region of gas within a distance 𝑑𝑑 ′ of
the wall of the container.
• For a molecule touching or colliding with a wall,
attractive forces between molecules would result
in a force directed toward other molecules in the
container and away from the wall.
Real Gases
• B. Finite Range of Intermolecular Interaction:
• Let 𝑑𝑑 ′ = range of the intermolecular force.
• Δp = ∫ 𝐹𝐹⃗ 𝑑𝑑𝑑𝑑= impulse of force acting
on the molecule during collision.
• Real gas molecules feel
attractive force on them besides
the force exerted by
the wall.
Real Gases
• B. Finite Range of Intermolecular Interaction:
• A molecule touching the wall, experiences
forces from the attraction of other nearby
molecules lying within a hemisphere of radius
𝑑𝑑 ′ .
• Resultant intermolecular attraction acts away
from the wall.
• This will result in a smaller force on the wall
and hence a reduction of pressure.
Real Gases
• B. Finite Range of Intermolecular Interaction:
• Net inward attractive force on one molecule
∝ Number of molecules within hemisphere of
′ 𝑁𝑁
radius 𝑑𝑑 ∝ density of molecules = ,
𝑉𝑉
where 𝑁𝑁 = total number of molecules
• Total number of such molecules within a
surface layer of thickness 𝑑𝑑 ′ and on unit area
of the wall (used to calculate the pressure) is
again ∝ 𝑁𝑁/𝑉𝑉
Real Gases
• B. Finite Range of Intermolecular Interaction:
• Hence, the net reduction of force acting on unit
surface area of the wall due to collision of the
molecules influenced by attractive forces of other
molecules
Δ𝐹𝐹 ∝ Δ𝐹𝐹1−𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚 × No. of molecules colliding
𝑁𝑁 𝑁𝑁 Δ𝐹𝐹
Δ𝐹𝐹 ∝ × ⇒ Δ𝑃𝑃 = ∝ 𝑁𝑁/𝑉𝑉 2
𝑉𝑉 𝑉𝑉 𝐴𝐴
• The net effect of attractive intermolecular force is
to introduce a correction in pressure,
proportional to 𝑁𝑁/𝑉𝑉 2
Real Gases
• B. Finite Range of Intermolecular Interaction:
• We can equivalently write the correction
terms in terms of number of moles 𝑛𝑛 as :
−𝑎𝑎 𝑛𝑛/𝑉𝑉 2
since 𝑁𝑁 = 𝑛𝑛𝑁𝑁𝐴𝐴 .
• Van der Waals Equation of State: Hence, van
der Waals equation of state becomes:
𝑛𝑛𝑛𝑛𝑛𝑛 𝑛𝑛 2
𝑃𝑃 = − 𝑎𝑎
𝑉𝑉 − 𝑛𝑛𝑛𝑛 𝑉𝑉
Van der Waals Equation of State
• This can be rewritten as:
𝑛𝑛2
𝑃𝑃 + 𝑎𝑎 2 𝑉𝑉 − 𝑛𝑛𝑛𝑛 = 𝑛𝑛𝑛𝑛𝑛𝑛
𝑉𝑉
where 𝑎𝑎 and 𝑏𝑏 are constants specific to each
real gas.
• In terms of molar volume 𝑣𝑣 = 𝑉𝑉/𝑛𝑛, we get
𝑎𝑎
𝑃𝑃 + 2 𝑣𝑣 − 𝑏𝑏 = 𝑅𝑅𝑅𝑅
𝑣𝑣
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Real Gases
• Real gases are gases, under conditions in
which they do not obey the ideal gas law.
• It was observed experimentally that, for some
gases (e.g., CO2) at high pressures or low
temperatures, noticeable deviations from
ideal gas behavior were exhibited.
• A satisfying explanation of these deviations in
terms of intermolecular forces was first given
by van der Waals in 1873.
Real Gases
• Real gases do not obey many of the
assumptions of the kinetic theory of gases.
• A. Molecules have a finite, small but certain
nonzero volume.
• B. The range of force between the molecules
extend beyond “collision diameter” i.e forces
are not just contact force.
• In developing a modified equation of state, we
take into account the above facts.
Real Gases
• A. Finite Volume of Molecules: Consider each
molecule to be a sphere of radius 𝑟𝑟.
• Two molecules can not come closer together than
𝑑𝑑 = 2𝑟𝑟 (i.e. their separation can not be < 2𝑟𝑟).
• Around each molecule, there is an excluded
volume i.e. a region where another molecule
cannot penetrate/come.
• Excluded volume around one molecule = sphere
4
of radius 𝑑𝑑 = Ω = 𝜋𝜋𝑑𝑑 3 = 8(4𝜋𝜋𝑟𝑟 3 /3)
3
Real Gases
• A. Finite Volume of Molecules:
• But this is the excluded volume between a pair of
molecules.
• Hence, we must divide by two to avoid over-counting.
• Hence, excluded volume
′ Ω 4
for each molecule 𝑏𝑏 = = 4( 𝜋𝜋𝑟𝑟 3 )
2 3
• For one mole of molecules,
total excluded volume is
𝑏𝑏 = 𝑁𝑁𝐴𝐴 𝑏𝑏′
Note: Two molecules approaching each other the volume in
which they interact is a hemisphere not a full sphere. Hence the factor of ½
Real Gases
• A. Finite Volume of Molecules: Hence, the
excluded volume should be subtracted from the
actual volume of the container.
• The net volume is the free volume in which the
molecules could undergo translational motion.
• The net free volume = 𝑉𝑉 − 𝑛𝑛𝑛𝑛, where 𝑛𝑛 = no. of
moles of gas and 𝑏𝑏 =excluded volume per mole.
• Hence, the finite volume of gas molecule
introduces a corrected equation of state:
𝑛𝑛𝑛𝑛𝑛𝑛
𝑃𝑃 =
𝑉𝑉 − 𝑛𝑛𝑛𝑛
Real Gases
• A. Finite Volume of Molecules:
• Hence, the pressure of the real gas is greater
relative to that of an ideal gas, under the same
conditions.
• The reduced volume available to the gas
molecules implies that they make more
collisions with the walls and thereby increase
the pressure.
Real Gases
• B. Finite Range of Intermolecular Interaction:
• The attractive forces between molecules are
weaker, but have a much longer range than that
of the repulsive forces (i.e. of hard core radius).
• Consider a region of gas within a distance 𝑑𝑑 ′ of
the wall of the container.
• For a molecule touching or colliding with a wall,
attractive forces between molecules would result
in a force directed toward other molecules in the
container and away from the wall.
Real Gases
• B. Finite Range of Intermolecular Interaction:
• Let 𝑑𝑑 ′ = range of the intermolecular force.
• Δp = ∫ 𝐹𝐹⃗ 𝑑𝑑𝑑𝑑= impulse of force acting
on the molecule during collision.
• Real gas molecules feel
attractive force on them besides
the force exerted by
the wall.
Real Gases
• B. Finite Range of Intermolecular Interaction:
• A molecule touching the wall, experiences
forces from the attraction of other nearby
molecules lying within a hemisphere of radius
𝑑𝑑 ′ .
• Resultant intermolecular attraction acts away
from the wall.
• This will result in a smaller force on the wall
and hence a reduction of pressure.
Real Gases
• B. Finite Range of Intermolecular Interaction:
• Net inward attractive force on one molecule
∝ Number of molecules within hemisphere of
′ 𝑁𝑁
radius 𝑑𝑑 ∝ density of molecules = ,
𝑉𝑉
where 𝑁𝑁 = total number of molecules
• Total number of such molecules within a
surface layer of thickness 𝑑𝑑′ and on unit area
of the wall (used to calculate the pressure) is
again ∝ 𝑁𝑁/𝑉𝑉
Real Gases
• B. Finite Range of Intermolecular Interaction:
• Hence, the net reduction of force acting on unit
surface area of the wall due to collision of the
molecules influenced by attractive forces of other
molecules
Δ𝐹𝐹 ∝ Δ𝐹𝐹1−𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚 × No. of molecules colliding
𝑁𝑁 𝑁𝑁 Δ𝐹𝐹
Δ𝐹𝐹 ∝ × ⇒ Δ𝑃𝑃 = ∝ 𝑁𝑁/𝑉𝑉 2
𝑉𝑉 𝑉𝑉 𝐴𝐴
• The net effect of attractive intermolecular force is
to introduce a correction in pressure,
proportional to 𝑁𝑁/𝑉𝑉 2
Real Gases
• B. Finite Range of Intermolecular Interaction:
• We can equivalently write the correction
terms in terms of number of moles 𝑛𝑛 as :
−𝑎𝑎 𝑛𝑛/𝑉𝑉 2
since 𝑁𝑁 = 𝑛𝑛𝑁𝑁𝐴𝐴 .
• Van der Waals Equation of State: Hence, van
der Waals equation of state becomes:
𝑛𝑛𝑛𝑛𝑛𝑛 𝑛𝑛 2
𝑃𝑃 = − 𝑎𝑎
𝑉𝑉 − 𝑛𝑛𝑛𝑛 𝑉𝑉
Van der Waals Equation of State
• This can be rewritten as:
𝑛𝑛2
𝑃𝑃 + 𝑎𝑎 2 𝑉𝑉 − 𝑛𝑛𝑛𝑛 = 𝑛𝑛𝑛𝑛𝑛𝑛
𝑉𝑉
where 𝑎𝑎 and 𝑏𝑏 are constants specific to each
real gas.
• In terms of molar volume 𝑣𝑣 = 𝑉𝑉/𝑛𝑛, we get:
𝑎𝑎
𝑃𝑃 + 2 𝑣𝑣 − 𝑏𝑏 = 𝑅𝑅𝑅𝑅
𝑣𝑣
Van der Waals Gas
• Isotherms of van der Waals Gas: The
iostherms of CO2 may be well described by
the van der Waals equation of state.

Van der Waals isotherms


Van der Waals Gas
• The van der Waals isotherms in the 𝑃𝑃𝑃𝑃-diagram
generally show two extrema, one maximum and
one minimum.
• At the extremum points 𝜕𝜕𝜕𝜕/𝜕𝜕𝜕𝜕 𝑇𝑇 = 0
• At certain temperature, the maximum and
minimum coalesce into a single inflection point
where the sign of the curvature changes
𝜕𝜕 2 𝑃𝑃
2
=0
𝜕𝜕𝑉𝑉 𝑇𝑇
Van der Waals Gas
• Critical Point: A critical point of a substance is the
point in the 𝑃𝑃𝑃𝑃-diagram where the sign of the
curvature of the isotherms changes. At the critical
point thus:
𝜕𝜕2 𝑃𝑃 𝜕𝜕𝜕𝜕
= 0 besides =0
𝜕𝜕𝑉𝑉 2 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇
• The values of the pressure, volume and
temperature at the critical point are called the
critical pressure, critical volume and critical
temperature, respectively.
Van der Waals Gas
• Example-1: Calculate the values of the critical
pressure, volume and temperature (i.e. 𝑃𝑃𝐶𝐶 , 𝑉𝑉𝐶𝐶
and 𝑇𝑇𝐶𝐶 ) of a van der Waals gas, in terms of the
gas constants 𝑎𝑎 and 𝑏𝑏.
• Solution: The van der Waals equation of state
𝑛𝑛2
𝑃𝑃 + 𝑎𝑎 2 𝑉𝑉 − 𝑛𝑛𝑛𝑛 = 𝑛𝑛𝑛𝑛𝑛𝑛, may be written as:
𝑉𝑉
𝑛𝑛𝑛𝑛𝑛𝑛 𝑎𝑎𝑛𝑛2
𝑃𝑃 = − 2
𝑉𝑉 − 𝑛𝑛𝑛𝑛 𝑉𝑉
where 𝑎𝑎 and 𝑏𝑏 are constants for a particular gas.
Van der Waals Gas
• Solution-1 (Contd.): Hence, we get:
𝜕𝜕𝜕𝜕 𝑛𝑛𝑛𝑛𝑇𝑇 2𝑎𝑎𝑛𝑛2
=− 2 + 3
𝜕𝜕𝜕𝜕 𝑇𝑇 𝑉𝑉−𝑛𝑛𝑛𝑛 𝑉𝑉
𝜕𝜕 2 𝑃𝑃 2𝑛𝑛𝑛𝑛𝑛𝑛 6𝑎𝑎𝑛𝑛2
= − 4
𝜕𝜕𝑉𝑉 2 𝑇𝑇
𝑉𝑉 − 𝑛𝑛𝑛𝑛 3 𝑉𝑉
• At the critical point, thus:
−𝑛𝑛𝑛𝑛𝑇𝑇𝐶𝐶 𝑉𝑉𝐶𝐶3 + 2𝑎𝑎𝑛𝑛2 𝑉𝑉𝐶𝐶 − 𝑛𝑛𝑛𝑛 2
=0
2𝑛𝑛𝑛𝑛𝑇𝑇𝐶𝐶 𝑉𝑉𝐶𝐶4 − 6𝑎𝑎𝑛𝑛2 𝑉𝑉𝐶𝐶 − 𝑛𝑛𝑛𝑛 3
=0
Van der Waals Gas
• Solution (Contd.): Hence we get:
𝑛𝑛𝑛𝑛𝑇𝑇𝐶𝐶 𝑉𝑉𝐶𝐶3 = 2𝑎𝑎𝑛𝑛2 𝑉𝑉𝐶𝐶 − 𝑛𝑛𝑛𝑛 2 ---(a)
2𝑛𝑛𝑛𝑛𝑇𝑇𝐶𝐶 𝑉𝑉𝐶𝐶4 = 6𝑎𝑎𝑛𝑛2 𝑉𝑉𝐶𝐶 − 𝑛𝑛𝑛𝑛 3 ---(b)
• From the above two: 2𝑉𝑉𝐶𝐶 = 3(𝑉𝑉𝐶𝐶 − 𝑛𝑛𝑛𝑛)
• Thus, 𝑉𝑉𝐶𝐶 = 3𝑛𝑛𝑛𝑛
• Putting this into (a) above,
4𝑛𝑛 2 𝑏𝑏 2 8𝑎𝑎
𝑛𝑛𝑛𝑛𝑇𝑇𝐶𝐶 = 2𝑎𝑎𝑛𝑛2 ⇒ 𝑇𝑇𝐶𝐶 =
27𝑛𝑛3 𝑏𝑏3 27𝑏𝑏𝑏𝑏
Van der Waals Gas
• Solution (Contd.): Hence from the van der
Waals equation of state, we get:

𝑛𝑛𝑛𝑛𝑇𝑇𝐶𝐶 𝑎𝑎𝑛𝑛2 𝑛𝑛𝑛𝑛8𝑎𝑎 𝑎𝑎𝑛𝑛2


𝑃𝑃𝐶𝐶 = − 2 = − 2 2
𝑉𝑉𝐶𝐶 − 𝑛𝑛𝑛𝑛 𝑉𝑉𝐶𝐶 27𝑏𝑏𝑏𝑏 2𝑛𝑛𝑛𝑛 9𝑛𝑛 𝑏𝑏
8𝑎𝑎 𝑎𝑎 𝑎𝑎
⇒ 𝑃𝑃𝐶𝐶 = 2
− 2=
54𝑏𝑏 9𝑏𝑏 27𝑏𝑏 2
Van der Waals Gas
• Example-2: Express the van der Waals gas
constants 𝑎𝑎 and 𝑏𝑏 , in terms of the critical
pressure and temperature at the critical point.
• Solution: We have for a van der Waals gas,
8𝑎𝑎 𝑎𝑎
𝑉𝑉𝐶𝐶 = 3𝑛𝑛𝑛𝑛, 𝑇𝑇𝐶𝐶 = and 𝑃𝑃𝐶𝐶 =
27𝑏𝑏𝑏𝑏 27𝑏𝑏2
where 𝑎𝑎 and 𝑏𝑏 are van der Waals gas constants.
• The critical temperature, 𝑇𝑇𝐶𝐶 , and the critical
pressure, 𝑃𝑃𝐶𝐶 , of a substance can be easily
determined experimentally.
Van der Waals Gas
• Solution-2 (contd.): This is because, at the critical
point, distinct liquid and gas phases co-exist.
𝑇𝑇𝐶𝐶 8𝑎𝑎 27𝑏𝑏2 8𝑏𝑏 𝑅𝑅𝑇𝑇𝐶𝐶
• Hence, = = ⇒ 𝑏𝑏 =
𝑃𝑃𝐶𝐶 27𝑏𝑏𝑏𝑏 𝑎𝑎 𝑅𝑅 8𝑃𝑃𝐶𝐶
27 27𝑏𝑏2 𝑇𝑇𝐶𝐶2 27R2Tc2
• Again, 𝑏𝑏𝑏𝑏𝑏𝑏𝐶𝐶 = 𝑎𝑎 =
8 64𝑃𝑃𝐶𝐶2 64Pc2
7
• For CO2 , we have 𝑃𝑃𝐶𝐶 = 0.75 × 10 𝑃𝑃𝑃𝑃, at
𝑇𝑇𝐶𝐶 = 304 𝐾𝐾
• Hence, 𝑎𝑎 = 0.364 𝐽𝐽. 𝑚𝑚3 /𝑚𝑚𝑚𝑚𝑙𝑙 2 , 𝑏𝑏 = 4.27 ×
10−5 𝑚𝑚3 /𝑚𝑚𝑚𝑚𝑚𝑚
Van der Waals Gas
• Example-3: Express van der Waals equation of
state in dimensionless reduced form.
• Solution: The van der Waals equation of state
may be written as:
𝑛𝑛𝑛𝑛𝑛𝑛 𝑎𝑎𝑛𝑛2
𝑃𝑃 = − 2
𝑉𝑉 − 𝑛𝑛𝑛𝑛 𝑉𝑉
where, 𝑎𝑎 and 𝑏𝑏 are van der Waals gas constants,
and 𝑛𝑛 is the number of moles.
• The gas constants 𝑎𝑎 and 𝑏𝑏 differ for different
gases.
Van der Waals Gas
• Solution-3 (contd.): In terms of the critical constants,
we define “reduced” quantities:
𝑃𝑃𝑅𝑅 = 𝑃𝑃/𝑃𝑃𝐶𝐶 , 𝑉𝑉𝑅𝑅 = 𝑉𝑉/𝑉𝑉𝐶𝐶 and 𝑇𝑇𝑅𝑅 = 𝑇𝑇/𝑇𝑇𝐶𝐶 .
where the critical constants are:
8𝑎𝑎 𝑎𝑎
𝑉𝑉𝐶𝐶 = 3𝑛𝑛𝑛𝑛, 𝑇𝑇𝐶𝐶 = and 𝑃𝑃𝐶𝐶 =
27𝑏𝑏𝑏𝑏 27𝑏𝑏2
• Then van der Waals equation of state becomes:
𝑛𝑛𝑛𝑛𝑛𝑛 𝑎𝑎𝑛𝑛2
𝑃𝑃 = − 2
𝑉𝑉 − 𝑛𝑛𝑛𝑛 𝑉𝑉
𝑛𝑛𝑛𝑛𝑇𝑇𝑅𝑅 𝑇𝑇𝐶𝐶 𝑎𝑎𝑛𝑛2
⇒ 𝑃𝑃𝑅𝑅 𝑃𝑃𝐶𝐶 = − 2 2
𝑉𝑉𝑅𝑅 𝑉𝑉𝐶𝐶 − 𝑛𝑛𝑛𝑛 𝑉𝑉𝑅𝑅 𝑉𝑉𝐶𝐶
Van der Waals Gas
• Solution-3 (contd.): Thus,
2 𝑛𝑛𝑛𝑛𝑇𝑇𝑅𝑅
8𝑎𝑎 2 2
27𝑏𝑏 27𝑏𝑏𝑏𝑏 𝑎𝑎𝑛𝑛 27𝑏𝑏
𝑃𝑃𝑅𝑅 = − 2 2 2
𝑎𝑎 (𝑉𝑉𝑅𝑅 3𝑛𝑛𝑛𝑛 − 𝑛𝑛𝑛𝑛) 𝑉𝑉𝑅𝑅 9𝑛𝑛 𝑏𝑏 𝑎𝑎
𝑛𝑛𝑇𝑇𝑅𝑅 𝑏𝑏𝑏𝑏𝑏 27𝑛𝑛2
⇒ 𝑃𝑃𝑅𝑅 = − 2 2
𝑎𝑎𝑎𝑎𝑎𝑎𝑎(𝑉𝑉𝑅𝑅 −1/3) 𝑉𝑉𝑅𝑅 9𝑛𝑛
8𝑇𝑇𝑅𝑅 27 8 1 3
⇒ 𝑃𝑃𝑅𝑅 = − 2 = 𝑇𝑇𝑅𝑅 − 2
(3𝑉𝑉𝑅𝑅 −1) 𝑉𝑉𝑅𝑅 9 3 𝑉𝑉 − 1 𝑉𝑉𝑅𝑅
𝑅𝑅 3
Van der Waals Gas
• Solution-3 (contd.): Hence, we get
3 1 8
𝑃𝑃𝑅𝑅 + 2 𝑉𝑉𝑅𝑅 − = 𝑇𝑇𝑅𝑅
𝑉𝑉𝑅𝑅 3 3
• This equation is invariant for all van der Waals gases; that
is, the same reduced form equation of state applies, no
matter what 𝑎𝑎 and 𝑏𝑏 may be for the particular gas, to all
van der Waals gases.
• We can check that the reduced equation of state
satisfies: 𝑃𝑃𝑅𝑅 𝑇𝑇𝑅𝑅 = 1, 𝑉𝑉𝑅𝑅 = 1 = 1
Van der Waals Gas
• Principle of Corresponding State:
• The van der Waals equation is called semi-
empirical in the sense that although it is based
on physical arguments, it contains two constants,
𝑎𝑎 and 𝑏𝑏 specific for each molecule, which must
be evaluated by comparison with experimental
data.
• We get reduced equation of state of van der
3 1 8
Waals gas as: 𝑃𝑃𝑅𝑅 + (𝑉𝑉𝑅𝑅 − ) = 𝑇𝑇𝑅𝑅
𝑉𝑉𝑅𝑅2 3 3
Van der Waals Gas
• In the reduced form of the equation, the
material constants 𝑎𝑎 and 𝑏𝑏 do not appear
explicitly.
• Thus when the van der Waals equation is in
the reduced form, all van der Waals gases,
may be considered to be in the same state,
when the values of 𝑃𝑃𝑅𝑅 , 𝑉𝑉𝑅𝑅 , and 𝑇𝑇𝑅𝑅 are the
same.
Van der Waals Gas
• This is true if pressure, volume and
temperature of each gas is measured in units
of its particular values of 𝑃𝑃𝐶𝐶 , 𝑉𝑉𝐶𝐶 , and 𝑇𝑇𝐶𝐶 (and
not in SI units or some other units).
• This is called the principle of corresponding
states.
• It is a principle of universal similarity
established first by van der Waals.
Van der Waals Gas
• Compression Factor or Compressibility Factor:
• The van der Waals equation of state in terms of molar
volume 𝑣𝑣 = 𝑉𝑉𝑚𝑚 = 𝑉𝑉/𝑛𝑛 is:
𝑛𝑛𝑛𝑛𝑛𝑛 𝑎𝑎𝑛𝑛2 𝑅𝑅𝑅𝑅 𝑎𝑎
𝑃𝑃 = − 2 = − 2
𝑉𝑉 − 𝑛𝑛𝑛𝑛 𝑉𝑉 𝑣𝑣 − 𝑏𝑏 𝑣𝑣
𝑃𝑃𝑃𝑃 𝑣𝑣 𝑎𝑎
• Hence, we get, = −
𝑅𝑅𝑅𝑅 𝑣𝑣−𝑏𝑏 𝑣𝑣𝑣𝑣𝑣𝑣
• We can expand the first term on the right hand side as:
2 3
𝑣𝑣 𝑣𝑣 𝑏𝑏 𝑏𝑏 𝑏𝑏
= =1+ + + +⋯
𝑣𝑣 − 𝑏𝑏 𝑣𝑣 1 − 𝑏𝑏 𝑣𝑣 𝑣𝑣 𝑣𝑣
𝑣𝑣
Van der Waals Gas
• Thus, we get
2 3
𝑃𝑃𝑃𝑃 𝑏𝑏 𝑏𝑏 𝑏𝑏 𝑎𝑎
=1+ + + + ⋯−
𝑅𝑅𝑅𝑅 𝑣𝑣 𝑣𝑣 𝑣𝑣 𝑣𝑣𝑣𝑣𝑣𝑣
1 𝑎𝑎 𝑏𝑏 3
= 1 + 𝑏𝑏 − + 𝑏𝑏 2 /𝑣𝑣 2 + 3 +⋯
𝑣𝑣 𝑅𝑅𝑅𝑅 𝑣𝑣
• Defining, 𝑍𝑍 = 𝑃𝑃𝑃𝑃/𝑅𝑅𝑅𝑅, we get for van der Waals gas:
1 𝑎𝑎 𝑏𝑏 3
𝑍𝑍vdW = 1 + 𝑏𝑏 − + 𝑏𝑏2 /𝑣𝑣 2 + 3 + ⋯
𝑣𝑣 𝑅𝑅𝑅𝑅 𝑣𝑣
• For ideal gas:
𝑃𝑃𝑃𝑃
𝑍𝑍𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖 = =1
𝑅𝑅𝑅𝑅
Van der Waals Gas
• This suggests that we can define the behaviour of
real gas in terms of the deviation of the
compressibility/compression factor 𝑍𝑍 from unity.
• Compression Factor: The compression factor is
defined as the ratio of the volume of the gas to
that of an ideal gas at the same temperature and
pressure:
𝑃𝑃𝑃𝑃 𝑃𝑃𝑃𝑃
𝑍𝑍 = = = 𝑣𝑣/𝑣𝑣𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑉𝑉/𝑉𝑉𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖
𝑛𝑛𝑛𝑛𝑛𝑛 𝑅𝑅𝑅𝑅
Van der Waals Gas
• At Intermediate Pressure: 𝒁𝒁 < 𝟏𝟏: Compression is
favored, due to dominance of attractive forces.
• At High Pressure: 𝒁𝒁 > 𝟏𝟏: Expansion is favored, as
repulsive forces come into play.
Van der Waals Gas
• Reduced Compressibility Factor: In the reduced form,
van der Waals equation of state can be written as:
3 1 8
𝑃𝑃𝑅𝑅 + 2 𝑉𝑉𝑅𝑅 − = 𝑇𝑇𝑅𝑅
𝑉𝑉𝑅𝑅 3 3
1 𝑇𝑇𝑅𝑅 𝛼𝛼
⇒ 𝑃𝑃𝑅𝑅 = − 2
𝑍𝑍𝐶𝐶 𝑉𝑉𝑅𝑅 − 𝛽𝛽 𝑉𝑉𝑅𝑅
where, 𝛽𝛽 = 1/3, 𝛼𝛼 = 9/8 and 𝑍𝑍𝐶𝐶 = 3/8,
• 𝑍𝑍𝐶𝐶 is the reduced compression factor at the critical
point (𝑉𝑉𝐶𝐶 = 3𝑛𝑛𝑛𝑛 = 𝑛𝑛𝑣𝑣𝐶𝐶 ):
𝑃𝑃𝐶𝐶 𝑣𝑣𝐶𝐶 𝑎𝑎 1 27𝑏𝑏𝑏𝑏 3
𝑍𝑍𝐶𝐶 = = 3𝑏𝑏 =
𝑅𝑅𝑇𝑇𝐶𝐶 27𝑏𝑏2 𝑅𝑅 8𝑎𝑎 8
Van der Waals Gas
• In terms of the reduced compression factor 𝑍𝑍𝐶𝐶 ,
the compression factor 𝑍𝑍 for van der Waals or
real gases becomes:
𝑃𝑃𝑃𝑃 𝑃𝑃𝑇𝑇𝐶𝐶 𝑣𝑣 𝑃𝑃𝐶𝐶 𝑣𝑣𝐶𝐶 3 𝑃𝑃𝑅𝑅 𝑣𝑣𝑅𝑅
𝑍𝑍 = = =
𝑅𝑅𝑅𝑅 𝑃𝑃𝐶𝐶 𝑇𝑇𝑣𝑣𝐶𝐶 𝑅𝑅𝑇𝑇𝐶𝐶 8 𝑇𝑇𝑅𝑅
• Hence, at a particular 𝑇𝑇𝑅𝑅 , the compression factor
will depend on 𝑃𝑃𝑅𝑅 and 𝑣𝑣𝑅𝑅 .
• Since, from the reduced equation of state, for
particular values of 𝑃𝑃𝑅𝑅 and 𝑇𝑇𝑅𝑅 , the reduced molar
volume 𝑣𝑣𝑅𝑅 = 𝑉𝑉𝑅𝑅 /𝑛𝑛 is the same for all real gases,
𝑍𝑍 will have the same behaviour for all real gases.
Van der Waals Gas
• In terms of the reduced pressure 𝑃𝑃𝑅𝑅 , the
compression factor behaves similarly for all
real gases (with different curves for different
𝑇𝑇
𝑇𝑇𝑅𝑅 = ): 𝑇𝑇𝑅𝑅 =2.0
𝑇𝑇𝐶𝐶
𝑇𝑇𝑅𝑅 =1.2

𝑇𝑇𝑅𝑅 =1.0

At different 𝑇𝑇𝑅𝑅
Van der Waals Gas
• Behaviour of compression factor:
Van der Waals Gas
• Virial Expansion for van der Waals gas (Virial
derived from the Latin word for ‘force’):
• For a real (van der Waals) gas, deviations from
ideal gas behavior result from intermolecular
forces and finite volume of the gas molecules.
• The intermolecular forces go to zero as the
average distance between molecules gets very
large.
• Hence, we expect that the compression factor 𝑍𝑍
would approach unity for real gases, as their
molar density, 1/𝑣𝑣 = 𝑛𝑛/𝑉𝑉 approaches zero.
Van der Waals Gas
• This suggests that it might be useful to expand 𝑍𝑍 in a
series of powers of the molar density:
𝑍𝑍 = 1 + 𝐵𝐵 1/𝑣𝑣 + 𝐶𝐶 1/𝑣𝑣 2 + ⋯
• We can also define the expansion of the compression
factor in terms of pressure as:
𝑍𝑍 = 1 + 𝐵𝐵′ 𝑃𝑃 + 𝐶𝐶 ′ 𝑃𝑃2 + ⋯
• These expansions are known as the virial expansions.
• The coefficients 𝐵𝐵, 𝐵𝐵′ , 𝐶𝐶, 𝐶𝐶 ′ , and so forth, which are
functions of temperature, are known as the virial
coefficients.
Van der Waals Gas
• Virial Expansions: The first virial coefficients
are unity.
• 𝐵𝐵(𝑇𝑇) and 𝐵𝐵′ (𝑇𝑇) are known as the second
virial coefficients.
• Although the virial expansions might seem
very complicated, because they contain an
infinite number of terms, their power lies in
the fact that, usually, only a few terms must
be considered.
Van der Waals Gas
• As pressure is reduced and molar volume gets very large, the
higher terms in the expansion become negligible and only the
first two terms need be considered:
𝑣𝑣𝑣𝑣 𝐵𝐵 𝑇𝑇
𝑍𝑍 ≈ =1+
𝑅𝑅𝑅𝑅 𝑣𝑣
𝑅𝑅𝑅𝑅 𝑅𝑅𝑅𝑅 𝑅𝑅𝑅𝑅 1
⇒ 𝑣𝑣 ≈ + 𝐵𝐵 𝑇𝑇 = + 𝐵𝐵 𝑇𝑇
𝑃𝑃 𝑃𝑃𝑃𝑃 𝑃𝑃 𝑍𝑍
𝑅𝑅𝑅𝑅
⇒ 𝑣𝑣 ≈ + 𝐵𝐵 𝑇𝑇
𝑃𝑃
• The final approximation results from 𝑣𝑣 being not too different
from the ideal gas molar volume (𝑅𝑅𝑅𝑅/𝑃𝑃) at low and moderate
pressures (to first order approximation, 𝑍𝑍 ≈ 1).
Van der Waals Gas
• Thus, the second virial coefficient 𝐵𝐵 𝑇𝑇 , is the
first correction to the ideal gas molar volume for
real gases.
• At very low temperature, molecules are strongly
influenced by attractive intermolecular forces (in
fact, if the temperature is low enough, the gas
will liquefy).
• This indicates that the second virial coefficient
should be negative at low temperatures (to make
𝑍𝑍 < 1).
Van der Waals Gas
• As the temperature is increased, thermal
velocities become so large that the weak
intermolecular attractive forces can have little
effect on molecular motions.
• The excluded volume, resulting from the
repulsive forces, should dominate under these
conditions, giving a positive second virial
coefficient and 𝑍𝑍 > 1 .
Van der Waals Gas
• The variation of second virial coefficient is
shown below:
Van der Waals Gas
• Boyle Temperature: The temperature at which
the second virial coefficient is zero is called
the Boyle temperature:
𝐵𝐵 𝑇𝑇𝐵𝐵 = 0
• At this temperature, attractive and repulsive
influences on the molar volume just cancel
and the gas behaves most ideally at low and
moderate pressures.
Quiz
• Boyle Temperature for the van der Waals gas:
Find an expression for the Boyle temperature of a
van der Waals gas.
• Solution: In order to set the second virial
coefficient equal to zero, we must put the van der
Waals equation into the virial form:
𝑍𝑍𝑅𝑅 = 1 + 𝐵𝐵 1/𝑣𝑣 + 𝐶𝐶 1/𝑣𝑣 2 + ⋯
• The second term in the van der Waals equation,
being proportional to (1/𝑣𝑣 2 ), is already in this
form.
Quiz
• In order to put the term 𝑅𝑅𝑅𝑅/(𝑣𝑣 − 𝑏𝑏) into the
virial form, we write it as:
𝑅𝑅𝑅𝑅 𝑅𝑅𝑅𝑅 1
=
𝑣𝑣 − 𝑏𝑏 𝑣𝑣 1 − 𝑏𝑏
𝑣𝑣
• Because 𝑏𝑏, the excluded volume per mole (the
actual size of the molecules) is much less than 𝑣𝑣,
𝑏𝑏
i.e. 𝑥𝑥 ≡ ≪ 1 we get,
𝑣𝑣
1
= 1 + 𝑥𝑥 + 𝑥𝑥 2 + 𝑥𝑥 3 + ⋯
1−𝑥𝑥
Quiz
𝑅𝑅𝑅𝑅 𝑏𝑏 𝑏𝑏 2 𝑏𝑏 3
• 𝑃𝑃 = 1+ + + + ⋯ − 𝑎𝑎/𝑣𝑣 2
𝑣𝑣 𝑣𝑣 𝑣𝑣 𝑣𝑣
𝑃𝑃𝑃𝑃 𝑎𝑎 1
• Or, 𝑍𝑍𝑅𝑅 = = 1 + 𝑏𝑏 − +⋯
𝑅𝑅𝑅𝑅 𝑅𝑅𝑅𝑅 𝑣𝑣
𝑎𝑎
• Hence, 𝐵𝐵 = 𝑏𝑏 −
𝑅𝑅𝑅𝑅
𝑎𝑎
• Hence the Boyle temperature is, 𝑇𝑇𝐵𝐵 = .
𝑅𝑅𝑅𝑅
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Van der Waals Gas
• Example-1: Calculate the values of the critical
pressure, volume and temperature (i.e. 𝑃𝑃𝐶𝐶 , 𝑉𝑉𝐶𝐶
and 𝑇𝑇𝐶𝐶 ) of a van der Waals gas, in terms of the
gas constants 𝑎𝑎 and 𝑏𝑏.
• Solution-1: The van der Waals equation of state
𝑛𝑛2
𝑃𝑃 + 𝑎𝑎 2 𝑉𝑉 − 𝑛𝑛𝑛𝑛 = 𝑛𝑛𝑛𝑛𝑛𝑛, may be written as:
𝑉𝑉
𝑛𝑛𝑛𝑛𝑛𝑛 𝑎𝑎𝑛𝑛2
𝑃𝑃 = − 2
𝑉𝑉 − 𝑛𝑛𝑛𝑛 𝑉𝑉
where 𝑎𝑎 and 𝑏𝑏 are constants for a particular gas.
Van der Waals Gas
• Solution-1 (contd.): Hence, we get:
𝜕𝜕𝜕𝜕 𝑛𝑛𝑛𝑛𝑇𝑇 2𝑎𝑎𝑛𝑛2
• =− + and
𝜕𝜕𝜕𝜕 𝑇𝑇 𝑉𝑉−𝑛𝑛𝑛𝑛 2 𝑉𝑉 3
𝜕𝜕2 𝑃𝑃 2𝑛𝑛𝑛𝑛𝑛𝑛 6𝑎𝑎𝑛𝑛2
• = −
𝜕𝜕𝑉𝑉 2 𝑇𝑇 𝑉𝑉−𝑛𝑛𝑛𝑛 3 𝑉𝑉 4
• At the critical point, 𝜕𝜕𝜕𝜕/𝜕𝜕𝜕𝜕 𝑇𝑇 = 0 and
𝜕𝜕 2 𝑃𝑃/𝜕𝜕𝑉𝑉 2 𝑇𝑇 = 0 :
−𝑛𝑛𝑛𝑛𝑇𝑇𝐶𝐶 𝑉𝑉𝐶𝐶3 + 2𝑎𝑎𝑛𝑛2 𝑉𝑉𝐶𝐶 − 𝑛𝑛𝑛𝑛 2 = 0
2𝑛𝑛𝑛𝑛𝑇𝑇𝐶𝐶 𝑉𝑉𝐶𝐶4 − 6𝑎𝑎𝑛𝑛2 𝑉𝑉𝐶𝐶 − 𝑛𝑛𝑛𝑛 3 = 0
Van der Waals Gas
• Solution-1 (contd.): Hence we get:
• 𝑛𝑛𝑛𝑛𝑇𝑇𝐶𝐶 𝑉𝑉𝐶𝐶3 = 2𝑎𝑎𝑛𝑛2 𝑉𝑉𝐶𝐶 − 𝑛𝑛𝑛𝑛 2 ---(a)
• 2𝑛𝑛𝑛𝑛𝑇𝑇𝐶𝐶 𝑉𝑉𝐶𝐶4 = 6𝑎𝑎𝑛𝑛2 𝑉𝑉𝐶𝐶 − 𝑛𝑛𝑛𝑛 3 ---(b)
• From the above two: 2𝑉𝑉𝐶𝐶 = 3(𝑉𝑉𝐶𝐶 − 𝑛𝑛𝑛𝑛)
• Thus, 𝑉𝑉𝐶𝐶 = 3𝑛𝑛𝑛𝑛
• Putting this into (a) above,
4𝑛𝑛 2 𝑏𝑏 2 8𝑎𝑎
𝑛𝑛𝑛𝑛𝑇𝑇𝐶𝐶 = 2𝑎𝑎𝑛𝑛2 ⇒ 𝑇𝑇𝐶𝐶 =
27𝑛𝑛3 𝑏𝑏3 27𝑏𝑏𝑏𝑏
Van der Waals Gas
• Solution-1 (contd.): Hence from the van der
Waals equation of state, we get:
𝑛𝑛𝑛𝑛𝑇𝑇𝐶𝐶 𝑎𝑎𝑛𝑛2 𝑛𝑛𝑛𝑛8𝑎𝑎 𝑎𝑎𝑛𝑛2
𝑃𝑃𝐶𝐶 = − 2 = − 2 2
𝑉𝑉𝐶𝐶 − 𝑛𝑛𝑛𝑛 𝑉𝑉𝐶𝐶 27𝑏𝑏𝑏𝑏 2𝑛𝑛𝑛𝑛 9𝑛𝑛 𝑏𝑏
8𝑎𝑎 𝑎𝑎 𝑎𝑎
⇒ 𝑃𝑃𝐶𝐶 = 2
− 2=
54𝑏𝑏 9𝑏𝑏 27𝑏𝑏 2
Van der Waals Gas
• Example-2: Express the van der Waals gas
constants 𝑎𝑎 and 𝑏𝑏 , in terms of the critical
pressure and temperature at the critical point.
• Solution-2: We have for a van der Waals gas,
8𝑎𝑎 𝑎𝑎
𝑉𝑉𝐶𝐶 = 3𝑛𝑛𝑛𝑛, 𝑇𝑇𝐶𝐶 = and 𝑃𝑃𝐶𝐶 =
27𝑏𝑏𝑏𝑏 27𝑏𝑏2
where 𝑎𝑎 and 𝑏𝑏 are van der Waals gas constants.
• The critical temperature, 𝑇𝑇𝐶𝐶 , and the critical
pressure, 𝑃𝑃𝐶𝐶 , of a substance can be easily
determined experimentally. In contrast, the
critical volume can not be determined easily.
Van der Waals Gas
• Solution-2 (contd.): This is because, at the critical
point, distinct liquid and gas phases co-exist.
𝑇𝑇𝐶𝐶 8𝑎𝑎 27𝑏𝑏2 8𝑏𝑏 𝑅𝑅𝑇𝑇𝐶𝐶
• Hence, = = ⇒ 𝑏𝑏 =
𝑃𝑃𝐶𝐶 27𝑏𝑏𝑏𝑏 𝑎𝑎 𝑅𝑅 8𝑃𝑃𝐶𝐶
27 27𝑅𝑅2 𝑇𝑇𝐶𝐶2
• Again, 𝑏𝑏𝑏𝑏𝑏𝑏𝐶𝐶 = 𝑎𝑎 =
8 64𝑃𝑃𝐶𝐶2
• For CO2 , we have 𝑃𝑃𝐶𝐶 = 0.75 × 107 𝑃𝑃𝑃𝑃 , at
𝑇𝑇𝐶𝐶 = 304 𝐾𝐾
• Hence, 𝑎𝑎 = 0.364 J.m3 /mol2 , 𝑏𝑏 = 4.27 ×
10−5 m3 /mol
Van der Waals Gas
• Example-3: Express van der Waals equation of
state in dimensionless reduced form.
• Solution:-3 The van der Waals equation of state
may be written as:
𝑛𝑛𝑛𝑛𝑛𝑛 𝑎𝑎𝑛𝑛2
𝑃𝑃 = − 2
𝑉𝑉 − 𝑛𝑛𝑛𝑛 𝑉𝑉
where, 𝑎𝑎 and 𝑏𝑏 are van der Waals gas constants,
and 𝑛𝑛 is the number of moles.
• The gas constants 𝑎𝑎 and 𝑏𝑏 differ for different
gases.
Van der Waals Gas
• Solution-3 (contd.): In terms of the critical constants,
we define “reduced” quantities:
𝑃𝑃𝑅𝑅 = 𝑃𝑃/𝑃𝑃𝐶𝐶 , 𝑉𝑉𝑅𝑅 = 𝑉𝑉/𝑉𝑉𝐶𝐶 and 𝑇𝑇𝑅𝑅 = 𝑇𝑇/𝑇𝑇𝐶𝐶 .
where the critical constants are:
8𝑎𝑎 𝑎𝑎
𝑉𝑉𝐶𝐶 = 3𝑛𝑛𝑛𝑛, 𝑇𝑇𝐶𝐶 = and 𝑃𝑃𝐶𝐶 =
27𝑏𝑏𝑏𝑏 27𝑏𝑏2
• Then van der Waals equation of state becomes:
𝑛𝑛𝑛𝑛𝑛𝑛 𝑎𝑎𝑛𝑛2
𝑃𝑃 = − 2
𝑉𝑉 − 𝑛𝑛𝑛𝑛 𝑉𝑉
𝑛𝑛𝑛𝑛𝑇𝑇𝑅𝑅 𝑇𝑇𝐶𝐶 𝑎𝑎𝑛𝑛2
⇒ 𝑃𝑃𝑅𝑅 𝑃𝑃𝐶𝐶 = − 2 2
𝑉𝑉𝑅𝑅 𝑉𝑉𝐶𝐶 − 𝑛𝑛𝑛𝑛 𝑉𝑉𝑅𝑅 𝑉𝑉𝐶𝐶
Van der Waals Gas
• Solution-3 (contd.): Thus,
2 𝑛𝑛𝑛𝑛𝑇𝑇𝑅𝑅
8𝑎𝑎 2 2
27𝑏𝑏 27𝑏𝑏𝑏𝑏 𝑎𝑎𝑛𝑛 27𝑏𝑏
𝑃𝑃𝑅𝑅 = − 2 2 2
𝑎𝑎 (𝑉𝑉𝑅𝑅 3𝑛𝑛𝑛𝑛 − 𝑛𝑛𝑛𝑛) 𝑉𝑉𝑅𝑅 9𝑛𝑛 𝑏𝑏 𝑎𝑎
𝑛𝑛𝑇𝑇𝑅𝑅 𝑏𝑏𝑏𝑏𝑏 27𝑛𝑛2
⇒ 𝑃𝑃𝑅𝑅 = − 2 2
𝑎𝑎𝑎𝑎𝑎𝑎𝑎(𝑉𝑉𝑅𝑅 −1/3) 𝑉𝑉𝑅𝑅 9𝑛𝑛
8𝑇𝑇𝑅𝑅 27 8 1 3
⇒ 𝑃𝑃𝑅𝑅 = − 2 = 𝑇𝑇𝑅𝑅 − 2
(3𝑉𝑉𝑅𝑅 −1) 𝑉𝑉𝑅𝑅 9 3 𝑉𝑉 − 1 𝑉𝑉𝑅𝑅
𝑅𝑅 3
Van der Waals Gas
• Solution-3 (contd.): Hence, we get
3 1 8
𝑃𝑃𝑅𝑅 + 2 𝑉𝑉𝑅𝑅 − = 𝑇𝑇𝑅𝑅
𝑉𝑉𝑅𝑅 3 3
• This equation is invariant for all van der Waals gases; that
is, the same reduced form equation of state applies, no
matter what 𝑎𝑎 and 𝑏𝑏 may be for the particular gas, to all
van der Waals gases.
• We can check that the reduced equation of state
satisfies: 𝑃𝑃𝑅𝑅 𝑇𝑇𝑅𝑅 = 1, 𝑉𝑉𝑅𝑅 = 1 = 1
Van der Waals Gas
• Compression Factor or Compressibility Factor:
• The van der Waals equation of state in terms of molar
volume 𝑣𝑣 = 𝑉𝑉𝑚𝑚 = 𝑉𝑉/𝑛𝑛 is:
𝑛𝑛𝑛𝑛𝑛𝑛 𝑎𝑎𝑛𝑛2 𝑅𝑅𝑅𝑅 𝑎𝑎
𝑃𝑃 = − 2 = − 2
𝑉𝑉 − 𝑛𝑛𝑛𝑛 𝑉𝑉 𝑣𝑣 − 𝑏𝑏 𝑣𝑣
𝑃𝑃𝑃𝑃 𝑣𝑣 𝑎𝑎
• Hence, we get, = −
𝑅𝑅𝑅𝑅 𝑣𝑣−𝑏𝑏 𝑣𝑣𝑣𝑣𝑣𝑣
• We can expand the first term on the right hand side as:
2 3
𝑣𝑣 𝑣𝑣 𝑏𝑏 𝑏𝑏 𝑏𝑏
= =1+ + + +⋯
𝑣𝑣 − 𝑏𝑏 𝑣𝑣 1 − 𝑏𝑏 𝑣𝑣 𝑣𝑣 𝑣𝑣
𝑣𝑣
Van der Waals Gas
• Thus, we get
2 3
𝑃𝑃𝑃𝑃 𝑏𝑏 𝑏𝑏 𝑏𝑏 𝑎𝑎
=1+ + + + ⋯−
𝑅𝑅𝑅𝑅 𝑣𝑣 𝑣𝑣 𝑣𝑣 𝑣𝑣𝑣𝑣𝑣𝑣
1 𝑎𝑎 𝑏𝑏 3
= 1 + 𝑏𝑏 − + 𝑏𝑏 2 /𝑣𝑣 2 + 3 +⋯
𝑣𝑣 𝑅𝑅𝑅𝑅 𝑣𝑣
• Defining, 𝑍𝑍 = 𝑃𝑃𝑃𝑃/𝑅𝑅𝑅𝑅, we get for van der Waals gas:
1 𝑎𝑎 𝑏𝑏 3
𝑍𝑍vdW = 1 + 𝑏𝑏 − + 𝑏𝑏2 /𝑣𝑣 2 + 3 + ⋯
𝑣𝑣 𝑅𝑅𝑅𝑅 𝑣𝑣
• For ideal gas:
𝑃𝑃𝑃𝑃
𝑍𝑍𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖 = =1
𝑅𝑅𝑅𝑅
Van der Waals Gas
• This suggests that we can define the behaviour of
real gas in terms of the deviation of the
compressibility/compression factor 𝑍𝑍 from unity.
• Compression Factor: The compression factor is
defined as the ratio of the volume of the gas to
that of an ideal gas at the same temperature and
pressure:
𝑃𝑃𝑃𝑃 𝑃𝑃𝑃𝑃
𝑍𝑍 = = = 𝑣𝑣/𝑣𝑣𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑉𝑉/𝑉𝑉𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖𝑖
𝑛𝑛𝑛𝑛𝑛𝑛 𝑅𝑅𝑅𝑅
Van der Waals Gas
• At Intermediate Pressure: 𝒁𝒁 < 𝟏𝟏: Compression is
favored, due to dominance of attractive forces.
• At High Pressure: 𝒁𝒁 > 𝟏𝟏: Expansion is favored, as
repulsive forces come into play.
Van der Waals Gas
• Reduced Compressibility Factor: In the reduced form,
van der Waals equation of state can be written as:
3 1 8
𝑃𝑃𝑅𝑅 + 2 𝑉𝑉𝑅𝑅 − = 𝑇𝑇𝑅𝑅
𝑉𝑉𝑅𝑅 3 3
1 𝑇𝑇𝑅𝑅 𝛼𝛼
⇒ 𝑃𝑃𝑅𝑅 = − 2
𝑍𝑍𝐶𝐶 𝑉𝑉𝑅𝑅 − 𝛽𝛽 𝑉𝑉𝑅𝑅
where, 𝛽𝛽 = 1/3, 𝛼𝛼 = 9/8 and 𝑍𝑍𝐶𝐶 = 3/8,
• 𝑍𝑍𝐶𝐶 is the reduced compression factor at the critical
point (𝑉𝑉𝐶𝐶 = 3𝑛𝑛𝑛𝑛 = 𝑛𝑛𝑣𝑣𝐶𝐶 ):
𝑃𝑃𝐶𝐶 𝑣𝑣𝐶𝐶 𝑎𝑎 1 27𝑏𝑏𝑏𝑏 3
𝑍𝑍𝐶𝐶 = = 3𝑏𝑏 =
𝑅𝑅𝑇𝑇𝐶𝐶 27𝑏𝑏2 𝑅𝑅 8𝑎𝑎 8
Van der Waals Gas
• In terms of the reduced compression factor 𝑍𝑍𝐶𝐶 ,
the compression factor 𝑍𝑍 for van der Waals or
real gases becomes:
𝑃𝑃𝑃𝑃 𝑃𝑃𝑇𝑇𝐶𝐶 𝑣𝑣 𝑃𝑃𝐶𝐶 𝑣𝑣𝐶𝐶 3 𝑃𝑃𝑅𝑅 𝑣𝑣𝑅𝑅
𝑍𝑍 = = =
𝑅𝑅𝑅𝑅 𝑃𝑃𝐶𝐶 𝑇𝑇𝑣𝑣𝐶𝐶 𝑅𝑅𝑇𝑇𝐶𝐶 8 𝑇𝑇𝑅𝑅
• Hence, at a particular 𝑇𝑇𝑅𝑅 , the compression factor
will depend on 𝑃𝑃𝑅𝑅 and 𝑣𝑣𝑅𝑅 .
• Since, from the reduced equation of state, for
particular values of 𝑃𝑃𝑅𝑅 and 𝑇𝑇𝑅𝑅 , the reduced molar
volume 𝑣𝑣𝑅𝑅 = 𝑉𝑉𝑅𝑅 /𝑛𝑛 is the same for all real gases,
𝑍𝑍 will have the same behaviour for all real gases.
Van der Waals Gas
• In terms of the reduced pressure 𝑃𝑃𝑅𝑅 , the
compression factor behaves similarly for all
real gases (with different curves for different
𝑇𝑇
𝑇𝑇𝑅𝑅 = ): 𝑇𝑇𝑅𝑅 =2.0
𝑇𝑇𝐶𝐶
𝑇𝑇𝑅𝑅 =1.2

𝑇𝑇𝑅𝑅 =1.0

At different 𝑇𝑇𝑅𝑅
Van der Waals Gas
• Behaviour of compression factor:
Van der Waals Gas
• Virial Expansion for van der Waals gas (Virial
derived from the Latin word for ‘force’):
• For a real (van der Waals) gas, deviations from
ideal gas behavior result from intermolecular
forces and finite volume of the gas molecules.
• The intermolecular forces go to zero as the
average distance between molecules gets very
large.
• Hence, we expect that the compression factor 𝑍𝑍
would approach unity for real gases, as their
molar density, 1/𝑣𝑣 = 𝑛𝑛/𝑉𝑉 approaches zero.
Van der Waals Gas
• This suggests that it might be useful to expand 𝑍𝑍 in a
series of powers of the molar density:
𝑍𝑍 = 1 + 𝐵𝐵 1/𝑣𝑣 + 𝐶𝐶 1/𝑣𝑣 2 + ⋯
• We can also define the expansion of the compression
factor in terms of pressure as:
𝑍𝑍 = 1 + 𝐵𝐵′ 𝑃𝑃 + 𝐶𝐶 ′ 𝑃𝑃2 + ⋯
• These expansions are known as the virial expansions.
• The coefficients 𝐵𝐵, 𝐵𝐵′ , 𝐶𝐶, 𝐶𝐶 ′ , and so forth, which are
functions of temperature, are known as the virial
coefficients.
Van der Waals Gas
• Virial Expansions: The first virial coefficients
are unity.
• 𝐵𝐵(𝑇𝑇) and 𝐵𝐵′ (𝑇𝑇) are known as the second
virial coefficients.
• Although the virial expansions might seem
very complicated, because they contain an
infinite number of terms, their power lies in
the fact that, usually, only a few terms must
be considered.
Van der Waals Gas
• As pressure is reduced and molar volume gets very large, the
higher terms in the expansion become negligible and only the
first two terms need be considered:
𝑣𝑣𝑣𝑣 𝐵𝐵 𝑇𝑇
𝑍𝑍 ≈ =1+
𝑅𝑅𝑅𝑅 𝑣𝑣
𝑅𝑅𝑅𝑅 𝑅𝑅𝑅𝑅 𝑅𝑅𝑅𝑅 1
⇒ 𝑣𝑣 ≈ + 𝐵𝐵 𝑇𝑇 = + 𝐵𝐵 𝑇𝑇
𝑃𝑃 𝑃𝑃𝑃𝑃 𝑃𝑃 𝑍𝑍
𝑅𝑅𝑅𝑅
⇒ 𝑣𝑣 ≈ + 𝐵𝐵 𝑇𝑇
𝑃𝑃
• The final approximation results from 𝑣𝑣 being not too different
from the ideal gas molar volume (𝑅𝑅𝑅𝑅/𝑃𝑃) at low and moderate
pressures (to first order approximation, 𝑍𝑍 ≈ 1).
Van der Waals Gas
• Thus, the second virial coefficient 𝐵𝐵 𝑇𝑇 , is the
first correction to the ideal gas molar volume for
real gases.
• At very low temperature, molecules are strongly
influenced by attractive intermolecular forces (in
fact, if the temperature is low enough, the gas
will liquefy).
• This indicates that the second virial coefficient
should be negative at low temperatures (to make
𝑍𝑍 < 1).
Van der Waals Gas
• As the temperature is increased, thermal
velocities become so large that the weak
intermolecular attractive forces can have little
effect on molecular motions.
• The excluded volume, resulting from the
repulsive forces, should dominate under these
conditions, giving a positive second virial
coefficient and 𝑍𝑍 > 1 .
Van der Waals Gas
• The variation of second virial coefficient is
shown below:
Van der Waals Gas
• Boyle Temperature: The temperature at which
the second virial coefficient is zero is called
the Boyle temperature:
𝐵𝐵 𝑇𝑇𝐵𝐵 = 0
• At this temperature, attractive and repulsive
influences on the molar volume just cancel
and the gas behaves most ideally at low and
moderate pressures.
Van der Waals Gas
• Example-4: Find the volume expansivity and
isothermal compressibility for a van der Waals
gas.
• Solution-4: The volume expansivity is defined as:
1 𝜕𝜕𝜕𝜕
𝛽𝛽 =
𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃
𝑛𝑛𝑛𝑛𝑛𝑛 𝑎𝑎𝑛𝑛2
• For a van der Waals gas: 𝑃𝑃 = −
𝑉𝑉−𝑛𝑛𝑛𝑛 𝑉𝑉 2
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
• Also from calculus: =−
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑉𝑉
Van der Waals Gas
𝜕𝜕𝜕𝜕
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑉𝑉
• Solution-4 (contd.): Hence, =− 𝜕𝜕𝜕𝜕
𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝑛𝑛𝑛𝑛
• For van der Waals gas: =
𝜕𝜕𝜕𝜕 𝑉𝑉 𝑉𝑉−𝑛𝑛𝑛𝑛
𝜕𝜕𝜕𝜕 −𝑛𝑛𝑛𝑛𝑛𝑛 𝑎𝑎𝑛𝑛2 −𝑛𝑛𝑛𝑛𝑛𝑛𝑉𝑉 3 +2𝑎𝑎𝑛𝑛2 𝑉𝑉−𝑛𝑛𝑛𝑛 2
• = + 2 3 =
𝜕𝜕𝜕𝜕 𝑇𝑇 𝑉𝑉−𝑛𝑛𝑛𝑛 2 𝑉𝑉 𝑉𝑉 3 𝑉𝑉−𝑛𝑛𝑛𝑛 2
1 𝜕𝜕𝜕𝜕 −1 𝑛𝑛𝑛𝑛 𝑉𝑉 3 𝑉𝑉−𝑛𝑛𝑛𝑛 2
• 𝛽𝛽 = =
𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃 𝑉𝑉 𝑉𝑉−𝑛𝑛𝑛𝑛 −𝑛𝑛𝑛𝑛𝑛𝑛𝑉𝑉 3 +2𝑎𝑎𝑛𝑛2 𝑉𝑉−𝑛𝑛𝑛𝑛 2
−𝑅𝑅𝑉𝑉 2 (𝑉𝑉−𝑛𝑛𝑛𝑛) 𝑅𝑅 𝑣𝑣−𝑏𝑏 𝑣𝑣 2
• ⇒ 𝛽𝛽 = =
2𝑎𝑎𝑎𝑎 𝑉𝑉−𝑛𝑛𝑛𝑛 2 −𝑅𝑅𝑅𝑅𝑉𝑉 3 𝑅𝑅𝑅𝑅𝑣𝑣 3 −2𝑎𝑎 𝑣𝑣−𝑏𝑏 2
Van der Waals Gas
1 𝜕𝜕𝜕𝜕
• Isothermal compressibility: 𝜅𝜅 = −
𝑉𝑉 𝜕𝜕𝜕𝜕 𝑇𝑇

1 1 𝑉𝑉 2 𝑉𝑉−𝑛𝑛𝑛𝑛 2
• 𝜅𝜅 = − 𝜕𝜕𝜕𝜕 =
𝑉𝑉 2𝑎𝑎𝑛𝑛2 𝑉𝑉−𝑛𝑛𝑛𝑛 2 −𝑛𝑛𝑛𝑛𝑛𝑛𝑉𝑉 3
𝜕𝜕𝜕𝜕 𝑇𝑇

𝑣𝑣 2 𝑣𝑣−𝑏𝑏 2
• 𝜅𝜅 =
2𝑎𝑎 𝑣𝑣−𝑏𝑏 2 −𝑅𝑅𝑅𝑅𝑣𝑣 3
Van der Waals Gas
• Example-5: The molar internal energy of a van
der Waals gas may be written as:
𝑎𝑎
𝑢𝑢 = 𝑐𝑐𝑐𝑐 −
𝑣𝑣
Find the molar specific heat 𝑐𝑐𝑉𝑉 of a van der
Waals gas.
• Solution-5: From the first law:
𝑑𝑑𝑑𝑑𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑑𝑑𝑄𝑄𝑖𝑖𝑖𝑖 + 𝑑𝑑𝑊𝑊𝑜𝑜𝑜𝑜
• For a gas, 𝑑𝑑𝑊𝑊𝑜𝑜𝑜𝑜 = −𝑃𝑃𝑃𝑃𝑃𝑃,
• At constant volume, 𝑑𝑑𝑑𝑑𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑑𝑑𝑄𝑄𝑖𝑖𝑖𝑖 = 𝑛𝑛𝑐𝑐𝑉𝑉 𝑑𝑑𝑑𝑑
Van der Waals Gas
1 𝑑𝑑 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖
• Solution-5 (contd.): Hence, 𝑐𝑐𝑉𝑉 =
𝑛𝑛 𝑑𝑑𝑑𝑑
1 𝑑𝑑 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 1 𝑑𝑑( 𝑛𝑛𝑛𝑛 ) 𝑑𝑑𝑑𝑑 𝑑𝑑 𝑎𝑎
• 𝑐𝑐𝑉𝑉 = = = = 𝑐𝑐𝑐𝑐 −
𝑛𝑛 𝑑𝑑𝑑𝑑 𝑛𝑛 𝑑𝑑𝑑𝑑 𝑑𝑑𝑑𝑑 𝑑𝑑𝑑𝑑 𝑣𝑣
• 𝑐𝑐𝑉𝑉 = 𝑐𝑐
Van der Waals Gas
• Example-6: 1 mol of a van der Waals gas undergoes a
reversible isothermal expansion from an initial molar
volume 𝑣𝑣𝑖𝑖 to a final molar volume 𝑣𝑣𝑓𝑓 . How much heat has
been transferred? (Given that, 𝜕𝜕𝜕𝜕/𝜕𝜕𝜕𝜕 𝑇𝑇 = 𝜕𝜕𝜕𝜕/𝜕𝜕𝜕𝜕 𝑉𝑉 )
• Solution-6: The entropy of a pure substance can be
considered as a function of any two variables, such as 𝑇𝑇
and 𝑉𝑉.
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
• Thus, 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑉𝑉
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝑉𝑉 𝑇𝑇
𝜕𝜕𝜕𝜕
• But, 𝑇𝑇𝑇𝑇𝑇𝑇 = 𝑑𝑑𝑑𝑑 = 𝑇𝑇 𝑑𝑑𝑑𝑑, for an isochoric process.
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕
• 𝐶𝐶𝑉𝑉 = 𝑛𝑛𝑐𝑐𝑉𝑉 = 𝑇𝑇
𝜕𝜕𝜕𝜕 𝑉𝑉
Van der Waals Gas
• Solution-6 (contd.): Again, it is given (and can be
shown) that:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
=
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕
• Hence, 𝑇𝑇𝑇𝑇𝑇𝑇 = 𝐶𝐶𝑉𝑉 𝑑𝑑𝑑𝑑 + 𝑇𝑇 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉
• In terms of the specific quantities:
𝜕𝜕𝜕𝜕
• 𝑇𝑇𝑇𝑇𝑇𝑇 = 𝑐𝑐𝑉𝑉 𝑑𝑑𝑑𝑑 + 𝑇𝑇 𝑑𝑑𝑣𝑣 where 𝑠𝑠, 𝑣𝑣, 𝑐𝑐𝑉𝑉 are molar
𝜕𝜕𝜕𝜕 𝑣𝑣
quantities.
𝑅𝑅𝑅𝑅 𝑎𝑎 𝜕𝜕𝜕𝜕 𝑅𝑅
• From 𝑃𝑃 = − we get, =
𝑣𝑣−𝑏𝑏 𝑣𝑣 2 𝜕𝜕𝜕𝜕 𝑣𝑣 𝑣𝑣−𝑏𝑏
Van der Waals Gas
• Solution-6 (contd.): Hence,
𝑑𝑑𝑑𝑑
𝑇𝑇𝑇𝑇𝑇𝑇 = 𝑐𝑐𝑉𝑉 𝑑𝑑𝑑𝑑 + 𝑅𝑅𝑅𝑅
𝑣𝑣 − 𝑏𝑏
• For constant 𝑇𝑇, 𝑑𝑑𝑑𝑑 = 0 ⇒ 𝑐𝑐𝑉𝑉 𝑑𝑑𝑑𝑑 = 0
• Since the process is reversible:
𝑞𝑞 = 𝑄𝑄 /𝑛𝑛 = ∫ 𝑑𝑑𝑑𝑑 /𝑛𝑛 = ∫ 𝑇𝑇𝑇𝑇𝑇𝑇
• Hence, total heat transferred in the reversible
isothermal process for van der Waals gas is:
𝑣𝑣𝑓𝑓 𝑑𝑑𝑑𝑑 𝑣𝑣𝑓𝑓 −𝑏𝑏
𝑞𝑞 = 𝑅𝑅𝑅𝑅 ∫𝑣𝑣 = 𝑅𝑅𝑅𝑅 ln
𝑖𝑖 𝑣𝑣−𝑏𝑏 𝑣𝑣𝑖𝑖 −𝑏𝑏
Van der Waals Gas
• Example-7: Show that the molar internal energy of a van
der Waals gas can be written as (given that 𝜕𝜕𝜕𝜕/𝜕𝜕𝜕𝜕 𝑇𝑇 =
𝑎𝑎
𝜕𝜕𝜕𝜕/𝜕𝜕𝜕𝜕 𝑉𝑉 ) : 𝑢𝑢 = ∫ 𝑐𝑐𝑉𝑉 𝑑𝑑𝑑𝑑 − 2 + const.
𝑣𝑣
• Strategy for the solution: The strategy for the solution of
the current problem should be:
• A. Since the problem asks for an expression for the internal
energy, we should use the first law.
• B. Since, it is given that 𝜕𝜕𝜕𝜕/𝜕𝜕𝜕𝜕 𝑇𝑇 = 𝜕𝜕𝜕𝜕/𝜕𝜕𝜕𝜕 𝑉𝑉 , entropy
must be used.
• C. Entropy is related to heat transferred 𝑑𝑑𝑑𝑑. Hence in the
first law we must use 𝑑𝑑𝑑𝑑 = 𝑇𝑇𝑇𝑇𝑇𝑇.
Van der Waals Gas
• Strategy for the solution:
• D. Since the problem asks for molar (or specific)
internal energy, we must use specific quantities.
• E. Since, the van der Waals equation of state has
𝑃𝑃 on the left hand side, taking derivative of 𝑃𝑃,
with respect to any other variable should be most
easy and we should express derivative of any
other quantity in terms of the derivative of 𝑃𝑃.
• Solution-7: From the first law:
𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑊𝑊𝑜𝑜𝑜𝑜
Van der Waals Gas
• Solution-7 (contd.):
𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖
• For a gas: 𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑇𝑇𝑇𝑇𝑇𝑇 − 𝑃𝑃𝑃𝑃𝑃𝑃 ⇒ =
𝑑𝑑𝑑𝑑
𝑑𝑑𝑑𝑑
𝑇𝑇 − 𝑃𝑃
𝑑𝑑𝑑𝑑
𝜕𝜕𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 𝜕𝜕𝜕𝜕
• Hence, = 𝑇𝑇 − 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇
• Using the given relation, we get:
𝜕𝜕𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 𝜕𝜕𝜕𝜕
= 𝑇𝑇 − 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉
Van der Waals Gas
• Solution-7 (contd.): For a van der Waals gas:
𝑅𝑅𝑅𝑅 𝑎𝑎 𝜕𝜕𝜕𝜕 𝑅𝑅
𝑃𝑃 = − ⇒ =
𝑣𝑣−𝑏𝑏 𝑣𝑣 2 𝜕𝜕𝜕𝜕 𝑣𝑣 𝑣𝑣−𝑏𝑏
𝜕𝜕𝜕𝜕 𝜕𝜕𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 𝑅𝑅
• Again, = = 𝑇𝑇 − 𝑃𝑃 = 𝑎𝑎/𝑣𝑣 2
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝑣𝑣−𝑏𝑏
𝑎𝑎
• ⇒ 𝑑𝑑𝑑𝑑 = 𝑐𝑐𝑉𝑉 𝑑𝑑𝑑𝑑 + 2 𝑑𝑑𝑑𝑑
𝑣𝑣
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 1 𝑑𝑑 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖
since, 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑 and 𝑐𝑐𝑉𝑉 =
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑣𝑣 𝑛𝑛 𝑑𝑑𝑑𝑑
𝑎𝑎
• 𝑢𝑢 = ∫ 𝑐𝑐𝑣𝑣 𝑑𝑑𝑑𝑑 − +const.
𝑣𝑣
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Simple Thermodynamic Systems
• A number of simple thermodynamic systems
include:
• A. Real Gas (e.g. van der Waals gas)
• B. Stretched Wire
• C. Thin Surfaces
• D. Dielectric Slab in a Capacitor
• E. Paramagnetic Rod inside a Magnetic Field
Simple Thermodynamic Systems
• B. Stretched Wire:
• A wire can be thought of as a one-dimensional
system, because its length is long compared
with its area.
• We assume:
• A. Extension of the wire at constant pressure
of standard atmospheric pressure.
• B. Change of volume as negligible.
Simple Thermodynamic Systems
• To calculate the volume change, we consider,
lateral change in radius= Δ𝑟𝑟 = 𝑟𝑟 ′ − 𝑟𝑟 < 0,
while the extension in length is: Δ𝐿𝐿 = 𝐿𝐿′ −
𝐿𝐿 > 0

𝝐𝝐𝒍𝒍𝒍𝒍𝒍𝒍
𝝈𝝈 = −
𝝐𝝐𝒍𝒍𝒍𝒍𝒍𝒍𝒍𝒍
Simple Thermodynamic Systems
• Define Poisson’s ratio 𝜎𝜎 as:
𝜎𝜎 = −(Δ𝑟𝑟/𝑟𝑟) Δ𝐿𝐿/ 𝐿𝐿 −1 = −𝜖𝜖𝑙𝑙𝑙𝑙𝑙𝑙 /𝜖𝜖𝑙𝑙𝑙𝑙𝑙𝑙𝑙𝑙
• Intrinsically, 0 < 𝜎𝜎. It can be shown that 0 < 𝜎𝜎 < 0.5.
• Hence, change in volume:
Δ𝑉𝑉 = 𝜋𝜋 𝑟𝑟 + Δ𝑟𝑟 2 𝐿𝐿 + Δ𝐿𝐿 − 𝐿𝐿𝑟𝑟 2
⇒ Δ𝑉𝑉 ≈ 𝜋𝜋 𝑟𝑟 2 Δ𝐿𝐿 + 2𝐿𝐿Δ𝑟𝑟 = 𝜋𝜋Δ𝐿𝐿𝑟𝑟 2 1 − 2𝜎𝜎
⇒ Δ𝑉𝑉/𝑉𝑉 = (1 − 2𝜎𝜎)Δ𝐿𝐿/𝐿𝐿 ⇒ 0 < Δ𝑉𝑉/𝑉𝑉 < Δ𝐿𝐿/(2𝐿𝐿)
• Since, Δ𝐿𝐿/𝐿𝐿 is a small quantity, so also Δ𝑉𝑉/𝑉𝑉.
• Hence, it is unnecessary to include the pressure and
the volume among the thermodynamic coordinates.
Simple Thermodynamic Systems
• A sufficiently complete thermodynamic description of
a wire is given in terms of only three coordinates:
A. The tension in the wire 𝐹𝐹, measured in newton (N).
B. The length of the wire L, measured in meters (m).
C. The absolute temperature T, measured in kelvin (K).
• Equation of State: For a wire at constant temperature
within the limits of elasticity, however, Hooke's law
holds:
𝐹𝐹 = −𝑘𝑘(𝐿𝐿 − 𝐿𝐿0 )
where 𝑘𝑘 is Hooke's constant and 𝐿𝐿0 is the length at zero
tension.
Simple Thermodynamic Systems
• If the tension is not balanced by an external force, the
wire will describe some sort of accelerated motion,
such as vibration, and will pass through states that
cannot be described in terms of thermodynamic
coordinates referring to the wire as a whole.
• The wire undergoes an infinitesimal change from one
state of equilibrium to another, then the infinitesimal
change of length is an exact differential and can be
written as:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝐹𝐹 𝜕𝜕𝜕𝜕 𝑇𝑇
Here, both the partial derivatives are functions of 𝑇𝑇 and 𝐹𝐹.
Simple Thermodynamic Systems
• Average Linear Co-efficient of Expansion: We define:
Average linear coefficient of expansion
change of length per unit length
=
change of temperature
at constant tension 𝐹𝐹.
• If the change of temperature becomes infinitesimal,
then the change of length also becomes infinitesimal.
• We define linear expansivity 𝛼𝛼, which is given by:
1 𝜕𝜕𝜕𝜕
𝛼𝛼 =
𝐿𝐿 𝜕𝜕𝜕𝜕 𝐹𝐹
Simple Thermodynamic Systems
• Since the length of a wire 𝐿𝐿 increases when the
temperature 𝑇𝑇 increases for metals, the linear
expansivity 𝛼𝛼 is most often a positive number for
metals.
• A rubber band, which may be considered a one-
dimensional system, has a negative linear
expansivity because it has a negative volume
expansivity.
• For small temperature range, 𝛼𝛼 may be regarded
as practically constant. It unit is 𝐾𝐾 −1
Simple Thermodynamic Systems
• Young’s Moldulus : We define
Average Young's modulus
change of tension per unit area
=
change of length per unit length
at constant temperature 𝑇𝑇.
• When the change of tension becomes infinitesimal, we
have what is known as the isothermal Young's modulus,
denoted by
𝐿𝐿 𝜕𝜕𝜕𝜕
𝑌𝑌 =
𝐴𝐴 𝜕𝜕𝜕𝜕 𝑇𝑇
where 𝐴𝐴 denotes the cross-sectional area.
Simple Thermodynamic Systems
• For small temperature range, 𝑌𝑌 may be
thought to be independent of 𝑇𝑇.
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
• Now, =−
𝜕𝜕𝜕𝜕 𝐿𝐿 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝐹𝐹
𝜕𝜕𝜕𝜕
• Hence, = −𝛼𝛼 𝐴𝐴 𝑌𝑌
𝜕𝜕𝜕𝜕 𝐿𝐿
• This means that for an increase in
temperature, the change of tension will be
opposite to the sign of the linear expansivity.
Simple Thermodynamic Systems
• C. Thin Surfaces:
• A surface can be considered as a two-dimensional
system. There are some important examples of
surfaces:
 (i) The interface of a liquid in equilibrium with a vapor.
 (ii) A soap bubble stretched across a wire framework,
consisting of two surfaces trapping a small amount of
liquid.
 (iii) A thin film of oil on the surface of water.
 (iv) A monolayer of a gas on the surface of a solid
catalyst.
Simple Thermodynamic Systems
• A thin surface, such as a soap film, is a
stretched membrane.
• The force/unit length,
acting perpendicularly
on a line drawn on the
membrane is called the
surface tension 𝛾𝛾 :
𝛾𝛾 = 𝐹𝐹/(2𝑙𝑙)
Simple Thermodynamic Systems
• An adequate thermodynamic description of a
surface is given by the specifying only three
coordinates:
• A. The surface tension 𝛾𝛾, measured in newton per
meter (N/m).
• B. The area of the film A, measured in square
meters (𝑚𝑚2 ).
• C. The absolute temperature 𝑇𝑇 , measured in
kelvin (K).
• Equations of State: Different thin surfaces have
different equations of state.
Simple Thermodynamic Systems
• (A) The surface of a pure liquid in equilibrium with its vapor
has an equation of state (developed by van der Waals and
Guggenheim):
𝑛𝑛
𝑇𝑇
𝛾𝛾 = 𝛾𝛾0 1 −
𝑇𝑇𝐶𝐶
where, 𝛾𝛾0 = surface tension at a standard temperature,
usually 20°C,
𝑇𝑇𝐶𝐶 = is the critical temperature (the temperature above
which no amount of pressure will condense vapor into liquid,
𝑛𝑛 = a constant that lying between 1 and 2.
• Thus the surface tension decreases as 𝑇𝑇 increases,
becoming zero when 𝑇𝑇 = 𝑇𝑇𝐶𝐶 .
Simple Thermodynamic Systems
• (B) The equation of state of a thin film of
insoluble oil on water is (within a restricted range
of values of area 𝐴𝐴):
𝛾𝛾 − 𝛾𝛾𝑊𝑊 𝐴𝐴 = 𝑎𝑎 𝑇𝑇
where, 𝛾𝛾𝑊𝑊 denotes the surface tension of a clean
water surface and
𝛾𝛾 = surface tension of the water covered by the
oil and 𝑎𝑎 is a constant.
• The difference 𝛾𝛾 − 𝛾𝛾𝑊𝑊 is sometimes called the
surface pressure or "two-dimensional" pressure.
Simple Thermodynamic Systems
• C. Dielectric Slab:
• Consider a capacitor consisting of two parallel
conducting plates of area 𝐴𝐴 with linear
dimensions that are large in comparison with
their separation 𝑙𝑙, filled with an isotropic solid
or liquid dielectric material.
Simple Thermodynamic Systems
• If a constant voltage is established across the plates,
then a uniform electric field 𝐸𝐸 exists in the dielectric
between the plates.
• The relation between the electric field, 𝐸𝐸 ,
displacement vector 𝐷𝐷 and the polarization vector 𝑃𝑃 is
given by:
𝐷𝐷 = 𝜖𝜖0 𝐸𝐸 + 𝑃𝑃
• In thermodynamics, it is not customary to refer to the
unit volume of a material, as 𝑃𝑃 does, because the
thermodynamic coordinate volume would be implicit
in the polarization.
Simple Thermodynamic Systems
• Rather, to show the explicit dependence on volume,
we introduce in thermodynamics the total polarization
Π, where 𝑃𝑃 = Π/𝑉𝑉 :
𝐷𝐷 = 𝜖𝜖0 𝐸𝐸 + Π/𝑉𝑉
• We may describe a dielectric with the aid of the three
thermodynamic coordinates:
 A. The electric field 𝐸𝐸, measured in volts per meter
(V/m).
 B. The total polarization Π ,measured in coulomb-
meters (C · m).
 C. The absolute temperature T, measured in kelvin (K).
Simple Thermodynamic Systems
• Equation of State: There are many dielectrics
whose equation of state at temperatures
above about 10 K is given by
Π 𝑏𝑏
= 𝑎𝑎 + 𝐸𝐸
𝑉𝑉 𝑇𝑇
where 𝑎𝑎, 𝑏𝑏 are constants.
Simple Thermodynamic Systems
• E. Paramagnetic Rod inside a Magnetic Field
• A paramagnetic material, has zero net magnetic
moment at zero applied field.
• Unlike a ferromagnet, such as iron, it has no hysteresis,
too.
• Upon being introduced into an external magnetic field,
a paramagnet becomes slightly magnetized in the
direction of the field.
• The atoms/molecules in a paramagnetic material has
unbalanced/unpaired electrons which have magnetic
moment due to spin angular momentum.
Simple Thermodynamic Systems
• This can be modelled by induced current
loops producing induced magnetic dipole
moments and net induced 𝐵𝐵𝑖𝑖𝑖𝑖𝑖𝑖 field.

𝑩𝑩 = 𝟎𝟎
Induced
Current loops
Net Induced
Current

𝑩𝑩 = 𝑩𝑩𝟎𝟎 + 𝑩𝑩𝒊𝒊𝒊𝒊𝒊𝒊
Simple Thermodynamic Systems
• The macroscopic effect of the induced
currents is called the magnetic dipole moment
per unit volume or the magnetization 𝑀𝑀 =
𝑑𝑑𝑚𝑚
.
𝑑𝑑𝑑𝑑
• The relation between the magnetic field
induction, 𝐵𝐵, magnetic field strength vector 𝐻𝐻
and the magnetization vector 𝑀𝑀 is given by:
𝐵𝐵 = 𝜇𝜇0 𝐻𝐻 + 𝑀𝑀
Simple Thermodynamic Systems
• In thermodynamics, it is not customary to
refer to the unit volume of a material, as 𝑀𝑀
does, because the thermodynamic coordinate
volume would be implicit in the
magnetization.
• To show the explicit dependence on volume,
we introduce in thermodynamics the total
magnetization 𝑋𝑋⃗, where 𝑀𝑀 = 𝑋𝑋⃗/𝑉𝑉.
Simple Thermodynamic Systems
• Then we get,
𝐵𝐵 = 𝜇𝜇0 𝐻𝐻 + 𝑋𝑋/𝑉𝑉
• We may describe a paramagnetic rod with the
aid of the three thermodynamic coordinates:
• A. The magnetic field strength 𝐻𝐻, measured in
amperes per meter (𝐴𝐴/𝑚𝑚).
• B. The total magnetization 𝑋𝑋 = 𝑀𝑀 𝑉𝑉 , measured
in ampere-meter squared (A.m2 ).
• C. The absolute temperature 𝑇𝑇, measured in
kelvin (K).
Simple Thermodynamic Systems
• Experiments shows that the magnetization of many
paramagnetic solids is a function of the ratio of the
magnetic intensity to the temperature.
• Equation of State: For a paramagnetic rod within a
magnetic field, the equation of state relating the
thermodynamic variables is given by the Curie’s law:
𝐻𝐻

𝑋𝑋 = 𝐶𝐶𝐶𝐶
𝑇𝑇
where, 𝐶𝐶𝑐𝑐 is called the Curie constant with units m3 .K
• Curie constant 𝐶𝐶𝑐𝑐 depends upon the amount of
material. For the Curie constant per mass, its units
become m3 .K/kg
Intensive and Extensive Coordinates
• Intensive and Extensive Coordinates :
• Imagine a system in equilibrium to be divided into
two equal parts, each with equal mass.
• Those properties of each half of the system that
remain the same are said to be intensive; those
that are halved are called extensive.
• The intensive coordinates of a system, such as
temperature and pressure, are independent of
the mass.
• The extensive properties are proportional to the
mass.
Intensive and Extensive Coordinates
• The thermodynamic coordinates below are
categorized as intensive and extensive:
Thermodynamic Intensive Coordinate Extensive Coordinate
Systems
Hydrostatic System Pressure 𝑃𝑃 Volume 𝑉𝑉
Stretched Wire Tension 𝐹𝐹 Length 𝐿𝐿
Thin Surface Surface tension 𝛾𝛾 Area 𝐴𝐴
Dielectric Slab Electric Field 𝐸𝐸 Total Polarization Π
Paramagnetic Rod Magnetic Field 𝐻𝐻 Total Magnetization 𝑋𝑋⃗
Intensive and Extensive Coordinates
• Extensive Properties/Coordinates: Properties of
thermodynamic systems that increase
proportionally with the size of the system are
called extensive coordinates. E.g. number of
moles 𝑛𝑛, volume 𝑉𝑉.
• Intensive properties/Coordinates: Properties of
thermodynamic systems that does not depend on
the size of the system and are defined for each
small region in the system are called intensive
coordinates. E.g. pressure 𝑃𝑃, temperature 𝑇𝑇.
Intensive and Extensive Coordinates
• Terms that are added together or are on
opposite sides of an equal sign must contain
the same number of extensive variables.
• The quotient or ratio of two extensive
variables is an intensive variable. For example,
we can define a molar volume, which is an
intensive coordinate by:
𝑉𝑉 𝑅𝑅𝑅𝑅
𝑣𝑣 = 𝑉𝑉𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚 = =
𝑛𝑛 𝑃𝑃
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Simple Thermodynamic Systems
• A number of simple thermodynamic systems
include:
• A. Real Gas
• B. Stretched Wire
• C. Thin Surfaces
• D. Dielectric Slab in a Capacitor
• E. Paramagnetic Rod inside a Magnetic Field
Simple Thermodynamic Systems
• E. Paramagnetic Rod inside a Magnetic Field
• A paramagnet, unlike a ferromagnet, such as iron,
has no permanent magnetic moment and no
hysteresis.
• Upon being introduced into an external magnetic
field, a paramagnet becomes slightly magnetized
in the direction of the field.
• This can be modelled by induced current loops
producing induced magnetic dipole moments.
Simple Thermodynamic Systems
• The macroscopic effect of the induced currents is
called the magnetic dipole moment per unit volume or
𝑑𝑑𝑚𝑚
the magnetization 𝑀𝑀 = .
𝑑𝑑𝑑𝑑
• The relation between the magnetic field induction, 𝐵𝐵,
magnetic field strength vector 𝐻𝐻 and the
magnetization vector 𝑀𝑀 is given by:
𝐵𝐵 = 𝜇𝜇0 𝐻𝐻 + 𝑀𝑀
• In thermodynamics, it is not customary to refer to the
unit volume of a material, as 𝑀𝑀 does, because the
thermodynamic coordinate volume would be implicit
in the magnetization.
Simple Thermodynamic Systems
• To show the explicit dependence on volume, we introduce
in thermodynamics the total magnetization 𝑋𝑋⃗, where
𝑀𝑀 = 𝑋𝑋⃗/𝑉𝑉.
• Then we get,
𝐵𝐵 = 𝜇𝜇0 𝐻𝐻 + 𝑋𝑋⃗/𝑉𝑉
• We may describe a paramagnetic rod with the aid of the
three thermodynamic coordinates:
• A. The magnetic field strength 𝐻𝐻, measured in amperes per
meter (𝐴𝐴/𝑚𝑚).
• B. The total magnetization 𝑋𝑋⃗ = 𝑀𝑀 𝑉𝑉 , measured in ampere-
meter squared (A.m2 ).
• C. The absolute temperature 𝑇𝑇, measured in kelvin (K).
Simple Thermodynamic Systems
• Experiments shows that the magnetization of many
paramagnetic solids is a function of the ratio of the
magnetic intensity to the temperature.
• Equation of State: For a paramagnetic rod within a
magnetic field, the equation of state relating the
thermodynamic variables is given by the Curie’s law:
𝐻𝐻

𝑋𝑋 = 𝐶𝐶𝐶𝐶
𝑇𝑇
where, 𝐶𝐶𝑐𝑐 is called the Curie constant with units m3 .K
• Curie constant 𝐶𝐶𝑐𝑐 depends upon the amount of
material. For the Curie constant per mass, its units
become m3 .K/kg
Work Done on Thermodynamic
Systems
• In thermodynamics, we deal with systems in
equilibrium or very nearly in equilibrium.
• Thermodynamic equilibrium implies all three
types of equilibrium:
Thermodynamic Equilibrium ≡
Mechanical Equilibrium + Chemical Equilibrium
+ Thermal Equilibrium
• When a finite unbalanced force is applied, the
system will go through a series of non-
equilibrium states.
Work Done on Thermodynamic
Systems
• Hence, during a process, if we desire to
describe every state of a system by means of
system-wide thermodynamic coordinates,
then the process must not be performed using
a finite unbalanced force or torque.
• Hence, in an ideal situation the external forces
acting on a system are varied only slightly.
• The unbalanced forces are infinitesimal, and
the process proceeds infinitesimally slowly.
Work Done on Thermodynamic
Systems
• Quasi-Static Processes : A thermodynamic
process in which the system is at all times
infinitesimally near a state of thermodynamic
equilibrium, and all the states through which
the system passes can be described by means
of thermodynamic coordinates referring to the
system as a whole, is called a quasi-static
process.
Quasi-static process≡ very slow process
Work Done on Thermodynamic
Systems
• In a quasi-static process:
• A. An equation of state hold all the times.
• B. There are no unbalanced forces or torques.
• C. All the changes are infinitesimal.
• D. External forces differ only by infinitesimal amount
from the internal forces.
• E. The system gets enough time to get into equilibrium
once an infinitesimal change has occurred.
• F. The thermodynamic variables are essentially the
same inside the system and at its boundary.
Work Done on Thermodynamic
Systems
• A quasi-static process is an idealization that is
applicable to all thermodynamics systems.
• The conditions for a quasi-static process can never be
rigorously satisfied in the laboratory, but they can be
approached with almost any degree of accuracy.
• In a finite quasi-static process in which the volume
changes from 𝑉𝑉𝑖𝑖 to 𝑉𝑉𝑓𝑓 , the amount of work 𝑊𝑊 done by
the system is:
𝑉𝑉𝐹𝐹
𝑊𝑊 = 𝑊𝑊𝑖𝑖𝑖𝑖 = − � 𝑃𝑃 𝑑𝑑𝑑𝑑
𝑉𝑉𝑖𝑖
Work Done on Thermodynamic
Systems
• A. Work in Changing the Length of a Wire:
• Let, 𝐿𝐿 = length of a wire, 𝐹𝐹 = a tension in the
wire and 𝑑𝑑𝑑𝑑 = the length of the wire that is
changed infinitesimally by an external force
opposite to the tension.
• Then, the infinitesimal amount of work that is
done on the wire is equal to
𝑑𝑑𝑑𝑑 = 𝐹𝐹 𝑑𝑑𝑑𝑑
where 𝐹𝐹 is a function of length 𝐿𝐿 and temperature
𝑇𝑇.
Work Done on Thermodynamic
Systems
• A positive value of 𝑑𝑑𝑑𝑑 i.e. 𝑑𝑑𝑑𝑑 > 0 means stretching for
which positive work must be done on the wire.
• For a finite change of length from 𝐿𝐿𝑖𝑖 to 𝐿𝐿𝑓𝑓 ,
𝐿𝐿𝑓𝑓
𝑊𝑊 = � 𝐹𝐹 𝑑𝑑𝑑𝑑
𝐿𝐿𝑖𝑖
• Note that, if the wire is undergoing a motion involving large
unbalanced forces, then the integral cannot be evaluated in
terms of thermodynamic coordinates referring to the wire
as a whole.
• If, however, the external force is maintained at all times
only slightly different from the tension, then the process is
sufficiently quasi-static and equation of state can be used.
Work Done on Thermodynamic
Systems
• B. Work In Changing the Area of a Surface Film:
• Consider a soap film, which consists of two
surfaces enclosing water, stretched across a wire
framework.
• The right side of the wire framework is movable
as shown :
• Let, 𝐿𝐿 = length of wire
• 𝛾𝛾 = surface tension of
soap solution.
• 𝐴𝐴 = area of the surface.
Work Done on Thermodynamic
Systems
• Then the external force 𝐹𝐹 exerted on both surfaces is
equal in magnitude to 2 𝛾𝛾𝛾𝛾
• For an infinitesimal displacement 𝑑𝑑𝑑𝑑, the work done is:
𝑑𝑑𝑑𝑑 = 2 𝛾𝛾𝛾𝛾 𝑑𝑑𝑑𝑑
• But for two films, 2 𝐿𝐿 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑
• Hence, 𝑑𝑑𝑑𝑑 = 𝛾𝛾𝛾𝛾𝛾𝛾
• For a finite change from an initial area 𝐴𝐴𝑖𝑖 to a final area
𝐴𝐴𝑓𝑓 , the work done is:
𝐴𝐴𝑓𝑓
𝑊𝑊 = � 𝛾𝛾 𝑑𝑑𝑑𝑑
𝐴𝐴𝑖𝑖
Work Done on Thermodynamic
Systems
• C. Work In Changing the Total Polarization of a
Dielectric Solid:
• Consider a slab of isotropic dielectric material
placed between the conducting plates of a
parallel-plate capacitor, as shown:

• Let, 𝐴𝐴 = the surface area of the capacitor plates.


Work Done on Thermodynamic
Systems
• Let, 𝑙𝑙 = separation between the plates ≪ 𝐴𝐴
• ℰ = constant potential difference equal to the
emf of a battery, if no current is drawn.
• The electric field 𝐸𝐸 will be nearly uniform
between the plates and given by 𝐸𝐸 = 𝐸𝐸 = ℰ/𝑙𝑙 .
• One plate is given a charge +𝑍𝑍 and the other a
charge −𝑍𝑍.
• When the charge of the capacitor is changed an
infinitesimal amount 𝑑𝑑𝑑𝑑, the work done is:
𝑑𝑑𝑑𝑑 = ℰ𝑑𝑑𝑑𝑑 = 𝐸𝐸𝐸𝐸 𝑑𝑑𝑑𝑑
Work Done on Thermodynamic
Systems
• The charge 𝑍𝑍 on the plates is equal to 𝑑𝑑𝑑𝑑 = 𝜎𝜎𝜎𝜎,
where 𝜎𝜎 = surface charge density.
• In a capacitor, at the edge of the dielectric, from
the boundary conditions, Δ𝐷𝐷𝑛𝑛 = 𝜎𝜎𝑓𝑓 .
• In side the metal, 𝐸𝐸 = 0 ⇒ 𝐷𝐷 = 0.
• Hence, Δ𝐷𝐷𝑛𝑛 = 𝐷𝐷𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠 − 𝐷𝐷𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚𝑚 = 𝐷𝐷𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠𝑠 = 𝜎𝜎𝑓𝑓
• Hence, the charge 𝑍𝑍 on the plates is equal to
𝑍𝑍 = 𝐷𝐷𝐷𝐷
Work Done on Thermodynamic
Systems
• Hence, work done is equal to
𝑑𝑑𝑑𝑑 = 𝐴𝐴𝐴𝐴𝐴𝐴 𝑑𝑑𝑑𝑑 = 𝑉𝑉𝑉𝑉 𝑑𝑑𝑑𝑑
where, 𝑉𝑉 =volume of the dielectric.
• If Π is the total polarization of the material
(assumed to be isotropic), then we have:
𝐷𝐷 = 𝜖𝜖0 𝐸𝐸 + Π/𝑉𝑉
• Thus the work done is:
𝑑𝑑𝑑𝑑 = 𝑉𝑉𝜖𝜖0 𝐸𝐸 𝑑𝑑𝑑𝑑 + 𝐸𝐸 𝑑𝑑Π
Work Done on Thermodynamic
Systems
• The first term is the work required to increase
the electric field by 𝑑𝑑𝑑𝑑 and would be present
even if a vacuum existed between the plates
of the capacitor.
• The second term is the work required to
increase the total polarization of the dielectric
𝑑𝑑Π.
• The second term is zero when no material is
present between the capacitor.
Work Done on Thermodynamic
Systems
• For the dielectric only, work done on the
dielectric is: 𝑑𝑑𝑑𝑑 = 𝐸𝐸 𝑑𝑑 Π
• If the total polarization is changed a finite
amount from 𝑃𝑃𝑖𝑖 to 𝑃𝑃𝑓𝑓 , then total work done
is:
Π𝑓𝑓
𝑊𝑊 = � 𝐸𝐸 𝑑𝑑Π
Π𝑖𝑖
Work Done on Thermodynamic
Systems
• D. Work in Changing the Total Magnetization of a
Paramagnetic Solid:
• Consider a sample of paramagnetic material in
the form of a ring of cross-sectional area 𝐴𝐴 and of
mean circumference 𝐿𝐿.
• Suppose that an insulated wire is wound around
the sample, forming a toroidal winding of 𝑁𝑁
closely spaced turns.
• A constant current may be maintained in the
winding by a battery and changed by means of
the variable resistor.
Work Done on Thermodynamic
Systems
• The direct current in the winding sets up a
magnetic induction 𝐵𝐵.
• 𝐵𝐵 will be nearly uniform over the cross-section of
the toroid.
• The current is changed
slowly.
• In time 𝑑𝑑𝑑𝑑 the magnetic
induction changes by an
amount 𝑑𝑑𝑑𝑑.
Work Done on Thermodynamic
Systems
• Faraday's principle of electromagnetic
induction, there is induced in the winding a
back emf ℰ, where
𝑑𝑑𝑑𝑑
ℰ = −𝑁𝑁𝑁𝑁
𝑑𝑑𝑑𝑑
• During the time interval 𝑑𝑑𝑑𝑑, a quantity of
charge 𝑑𝑑𝑑𝑑 moves in the circuit, and the work
done by the battery to maintain the current is:
𝑑𝑑𝑑𝑑 = −ℰ 𝑑𝑑𝑑𝑑
Work Done on Thermodynamic
Systems
• The work done by the battery is:
𝑑𝑑𝑑𝑑 𝑑𝑑𝑑𝑑
𝑑𝑑𝑑𝑑 = −ℰ 𝑑𝑑𝑑𝑑 = 𝑁𝑁𝑁𝑁 𝑑𝑑𝑑𝑑 = 𝑁𝑁𝑁𝑁 𝑑𝑑𝑑𝑑
𝑑𝑑𝑑𝑑 𝑑𝑑𝑑𝑑
Or, 𝑑𝑑𝑑𝑑 = 𝑁𝑁𝑁𝑁𝑁𝑁 𝑑𝑑𝑑𝑑
• The magnetic field 𝐻𝐻 due to a current 𝐼𝐼 inside a
toroidal winding is given:
𝑁𝑁𝑁𝑁 𝑁𝑁𝑁𝑁𝑁𝑁 𝑁𝑁𝑁𝑁𝑁𝑁
𝐻𝐻 = = =
𝐿𝐿 𝐴𝐴𝐴𝐴 𝑉𝑉
where 𝑉𝑉 is the volume of paramagnetic material.
Work Done on Thermodynamic
Systems
• Therefore, 𝑁𝑁𝑁𝑁𝑁𝑁 = 𝐻𝐻𝐻𝐻 and we get
𝑑𝑑𝑑𝑑 = 𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉
• If 𝑋𝑋⃗ = 𝑀𝑀𝑉𝑉 is the total magnetization of the
paramagnet (assumed to be isotropic), then
we have the relation: 𝐵𝐵 = 𝜇𝜇0 𝐻𝐻 + 𝑀𝑀
• Thus, 𝑑𝑑𝑑𝑑 = 𝜇𝜇0 𝑉𝑉𝑉𝑉 𝑑𝑑𝑑𝑑 + 𝜇𝜇0 𝐻𝐻 𝑑𝑑𝑑𝑑
• If no material were present within the
solenoidal winding, then 𝑋𝑋⃗ would be zero.
Work Done on Thermodynamic
Systems
• For no paramagnetic material:
𝑑𝑑𝑑𝑑 = 𝜇𝜇0 𝑉𝑉𝑉𝑉 𝑑𝑑𝑑𝑑
• This is the work necessary to increase the magnetic
field in a volume 𝑉𝑉 of empty space.
• The second term, 𝜇𝜇0 𝐻𝐻 𝑑𝑑𝑑𝑑 , is the work done in
increasing the total magnetization of the material by
an amount 𝑑𝑑𝑑𝑑: 𝑑𝑑𝑑𝑑 = 𝜇𝜇0 𝐻𝐻 𝑑𝑑𝑑𝑑
• If the total magnetization is caused to change a finite
amount from 𝑋𝑋𝑖𝑖 to 𝑋𝑋𝑓𝑓 , work done is:
𝑋𝑋𝑓𝑓
𝑊𝑊 = 𝜇𝜇0 � 𝐻𝐻 𝑑𝑑𝑑𝑑
𝑋𝑋𝑖𝑖
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Work Done on Thermodynamic
Systems
• In thermodynamics, we deal with systems in
equilibrium or very nearly in equilibrium.
• Thermodynamic equilibrium implies all three
types of equilibrium:
Thermodynamic Equilibrium ≡
Mechanical Equilibrium + Chemical Equilibrium
+ Thermal Equilibrium
• When a finite unbalanced force is applied, the
system will go through a series of non-
equilibrium states.
Generalized Force and Generalized
Work
• In the previous lectures we have studied five
types of simple thermodynamics systems:
• A. Real Gas
• B. Stretched Wire
• C. Thin Surfaces
• D. Dielectric Slab in a Capacitor
• E. Paramagnetic Rod inside a Magnetic Field
Generalized Force and Generalized
Work
• The work done on these systems are given in
the table below:

Thermodynamic Intensive Extensive Coordinate


Work Done 𝒅𝒅𝒅𝒅
Systems Coordinate
Hydrostatic System Pressure 𝑃𝑃 Volume 𝑉𝑉 −𝑃𝑃𝑃𝑃𝑃𝑃
Stretched Wire Tension 𝐹𝐹 Length 𝐿𝐿 𝐹𝐹𝐹𝐹𝐹𝐹
Thin Surface Surface tension 𝛾𝛾 Area 𝐴𝐴 𝛾𝛾𝛾𝛾𝛾𝛾
Dielectric Slab Electric Field 𝐸𝐸 Total Polarization Π 𝐸𝐸 𝑑𝑑Π
Paramagnetic Rod Magnetic Field 𝐻𝐻 Total Magnetization 𝑋𝑋⃗ 𝜇𝜇0 𝐻𝐻 𝑑𝑑𝑑𝑑
Generalized Force and Generalized
Work
• Note that, each expression of the work done is a
product of an intensive coordinate and an
extensive coordinate.
• Hence the work done is an extensive coordinate.
• A work diagram is obtained if anyone of the
intensive coordinates is plotted against its
corresponding extensive coordinate.
• There are, therefore, as many work diagrams as
there are systems.
Generalized Force and Generalized
Work
• Generalized Displacement: The extensive
thermodynamic variable whose change
multiplied by the appropriate generalized
conjugate variable, gives the work done on a
system is called the generalized displacement.
• In our systems these are: volume 𝑉𝑉, length 𝐿𝐿,
area 𝐴𝐴 , total polarization Π and the total
magnetization 𝑋𝑋.
• All these relate to the total size of the syste and
are extensive.
Generalized Force and Generalized
Work
• Generalized Force: The intensive thermodynamic
coordinates which multiplied with the
generalized displacements of the thermodynamic
systems gives the work done on the system are
called the generalized forces.
• For example, in our systems these are: pressure
𝑃𝑃, tension 𝐹𝐹, surface tension 𝛾𝛾, electric field 𝐸𝐸
and the magnetic intensity times permeability
𝜇𝜇0 𝐻𝐻.
Generalized Force and Generalized
Work
• In terms of the generalized variables, we may
represent the work done by any simple system
on a generalized work diagram by plotting the
generalized force Y against the generalized
displacement X.
• Conclusions based on such a diagram will hold
for any simple system.
Work and Heat
• A system could be transferred from an initial to a
final state by means of a quasi-static process.
• We have seen the expressions for the work done
on a system.
• There are, however, other means of changing the
state of a system that do not necessarily involve
the performance of work.
• Consider the following four processes which
involve closed systems (no matter or energy
passes between the system and its environment).
Work and Heat
• A. The system consists of water and a paddle wheel,
which is caused to rotate and churn the water by
means of a falling weight.
• As a result, the temperature of the water rises from
room temperature to a slightly higher temperature.
• B. In the second system the water and the resistor
constitute the system.
• The electric current in the resistor being maintained by
a generator turned by means of a falling weight.
• Again, the temperature of the water rises.
Work and Heat
• In both cases, the state of the system is
caused to change; and since the agency for
changing the state of the system is a falling
weight, both processes involve the
performance of work.
Work and Heat
• In the other two situations, system is water in a
diathermic container.
• In the third case system is in contact with the
burning gases at a high temperature; whereas, in
the fourth case, the system is near but not in
contact with a lamp whose temperature is much
higher than that of the water.
• In both cases, the system is caused to change, but
in neither case can the agency of change be
described by mechanical means.
Work and Heat
• In the last two cases, the state of the system
changes by means of flow of heat from one
part to the other.
Work and Heat
• We define heat as:
• Heat is that which is transferred between a
system and its surroundings by virtue of a
temperature difference only.
• We know that heat is a form of energy.
• We also define:
• Adiabatic Wall or heat insulator: A wall which is
impervious to heat transfer.
• Diathermic wall or heat conductor: A wall that
transmits heat.
Work and Heat
• Adiabatic Work: When a closed system is
completely surrounded by an adiabatic boundary,
the system may still be coupled to the
surroundings so that work may be done.
• Work done in an adiabatic container is called an
adiabatic work.
• The state of a system may be caused to change
from a given initial state to the same final state by
the performance of adiabatic work only.
Work and Heat
• We have different examples of adiabatic work.
• In all three cases, the system is surrounded by
adiabatic wall.
Work and Heat
• We may change the state of a system by different
adiabatic paths.
• There are an infinite
number of paths
by which a system
may be transferred
from an initial state
to a final state by
the performance of
adiabatic work only.
Work and Heat
• Experiment proved that the adiabatic work is the
same along all such paths.
• The generalization of the statement that the
adiabatic work is path independent is a restricted
statement of the first law of thermodynamics:

lf a closed system is caused to change from


an initial state to a final state by adiabatic
means only, then the work done on the system
is the same for all adiabatic paths connecting
the two states.
Work and Heat
• Whenever a quantity is known to depend only on the
initial and final states, and not on the path connecting
them, an important conclusion can be drawn.
• Therefore, it follows from the restricted statement of
the first law of thermodynamics that:

There exists a function of the coordinates of a


thermodynamic system whose value at the final state
minus its value at the initial state is equal to the
adiabatic work in going from one state to the other.
• This function is known as the internal-energy function.
Work and Heat
• Hence, we have
𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑈𝑈
𝑊𝑊𝑖𝑖→𝑓𝑓 adiabatic = 𝑈𝑈𝑓𝑓 − 𝑈𝑈𝑖𝑖
where, 𝑊𝑊𝑖𝑖→𝑓𝑓 denotes work done on the system.
• Thus, thermodynamic work, which is generally path-
dependent, becomes path-independent for an
adiabatic process.
• The equation above states that:
There exists an energy function, whose difference
between two values is the energy change of the system.
Work and Heat
• Internal Energy Function: For a hydrostatic
system, if U is regarded as a function of 𝑇𝑇 and 𝑉𝑉:
𝑈𝑈 = 𝑈𝑈 𝑇𝑇, 𝑉𝑉 and
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 𝑇𝑇, 𝑉𝑉 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑇𝑇
• Again, internal energy may be thought of a
function of 𝑇𝑇 and 𝑃𝑃 only: 𝑈𝑈 = 𝑈𝑈(𝑇𝑇, 𝑃𝑃)
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 𝑇𝑇, 𝑃𝑃 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑃𝑃
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝑃𝑃 𝑇𝑇
Work and Heat
• Mathematical Formulation of the First Law:
• The process of adiabatic work is not the only
process the state of a system may be changed.
• We may have processes involving diathermic wall
in which heat is exchanged between the
surroundings and the system as well as work
done.
• In the first case, the gas is in contact with a flame
whose temperature is higher than that of the gas.
Work and Heat
• At the same time, the gas is forced to contract, so
diathermic work is performed on the system.
• In the second case,
the total magnetization
of a paramagnetic solid
is increased while it
is in contact with liquid
helium.
Work and Heat
• The temperature of liquid helium is lower than
that of the solid.
• Hence, some of the helium boils away during the
magnetization.
• We may have work done on/by a system while
heat passing to/from the system.
• Consider two cases of same change of a closed
system:
• A. In the first system, only adiabatic work is done
to change the state of the system from 𝑖𝑖 to 𝑓𝑓 in
order to obtain 𝑈𝑈𝑖𝑖 − 𝑈𝑈𝑓𝑓 .
Work and Heat
• B. In the second case, we cause the system to
undergo the same change of state, so we have
the same 𝑈𝑈𝑖𝑖 − 𝑈𝑈𝑓𝑓 , but the process is diathermic.
• We measure the diathermic work 𝑊𝑊 done.
• The result of all such experiments is that the
nonadiabatic work 𝑊𝑊 is not equal to 𝑈𝑈𝑖𝑖 − 𝑈𝑈𝑓𝑓 .
• In order that this result shall be be consistent
with the law of the conservation of energy, we
are forced to conclude that energy has been
transferred by means other than the
performance of work.
Work and Heat
• This energy, whose transfer between the system and its
surroundings is required by the law of the conservation
of energy and which has taken place only by virtue of
the temperature difference between the system and its
surroundings, is called heat.
When a closed system whose surroundings are at a
different temperatures and on which diathermic work
may be done undergoes a process, then the energy
transferred by non-mechanical means, equal to the
difference between the change of internal energy and the
diathermic work, is called heat.
Work and Heat
• Mathematically, we define heat transferred as:
𝑄𝑄 = 𝑈𝑈𝑓𝑓 − 𝑈𝑈𝑖𝑖 − 𝑊𝑊(diathermic)
Or, 𝑈𝑈𝑓𝑓 − 𝑈𝑈𝑖𝑖 = 𝑄𝑄 + 𝑊𝑊
• Here, 𝑄𝑄 is the heat entering the system and 𝑊𝑊 is
the work done on the system.
• This is called the mathematical formulation of the
first law of thermodynamics. Recall our previous
notation:
Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄𝑖𝑖𝑖𝑖 + 𝑊𝑊𝑜𝑜𝑜𝑜
Work and Heat
• The mathematical formulation of the first law
contains three related ideas:
• (1) the existence of an internal-energy
function;
• (2) the principle of the conservation of energy;
• (3) the definition of heat as energy in transit
by virtue of a temperature difference.
Work and Heat
• Heating is a process by which there is an
exchange of energy between system and
surroundings because of a difference in
temperature.
• But what is the energy that is exchanged?
• The question cannot be answered until the
conditions for the process of heating are
determined.
• Consider a hydrostatic system and the following
two cases:
Work and Heat
• Case-A: The diathermic boundary of a hydrostatic
system is held rigid, then the volume of the
system does not change.
• Then the isochoric heat transferred is simply the
internal energy (work done is zero).
• Case-B: If the diathermic boundary of a
hydrostatic system is movable (a piston), then the
pressure of the system does not change.
• The isobaric heat transferred is known as the
enthalpy, which is another type of energy.
Work and Heat
• Concept of Heat: Heat is either internal energy or
enthalpy in transit, depending on the
experimental conditions.
• During the process of heating, energy flows from
one part of a system to another, or from one
system to another, by virtue of only a
temperature difference.
• When the flow has ceased, there is no longer any
occasion to use the word heat or the symbol Q,
because the process is completed.
Work and Heat
• All that remains after heating has been
completed is a different state of the system, that
is, a new value for the internal energy or
enthalpy.
• Consequently, it is incorrect to refer to the "heat
in a body," just as it is incorrect to speak of the
"work in a body."
• The processes of working and heating are
transient activities that lead to a change of the
energy found in a system.
• All that endures is the new state of the energy.
Work and Heat
• The energy of a system cannot be separated into
a mechanical part and a thermal part.
• This is analogous to water in a lake as originating
from this river, and other water from that rain.
The river and the rain have lost their meanings,
but the new water level endures.
• The work done on or by a system is not a function
of the coordinates of the system, so the
calculation of the work depends on the path of
integration by which the system is brought from
the initial to the final state.
Work and Heat
• The same situation applies to the heat
transferred in or out of a system.
• Heat 𝑄𝑄 is not a function of the thermodynamic
coordinates, that is, not a state function.
• So the calculation of the heat depends on the
path of integration.
• An infinitesimal amount of heat 𝑑𝑑𝑑𝑑, therefore, is
an inexact differential and not the differential of
an actual function of the thermodynamic
coordinates.
Work and Heat
• Equations of a Hydrostatic System:
• From the infinitesimal form of the first law:
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
⇒ 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 − 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑃𝑃𝑃𝑃𝑃𝑃
• Here, 𝑈𝑈 is a function of any two of 𝑃𝑃, 𝑉𝑉 and 𝑇𝑇.
• Using 𝑈𝑈 = 𝑈𝑈(𝑇𝑇, 𝑉𝑉) we get
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 𝑇𝑇, 𝑉𝑉 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑇𝑇
Work and Heat
• Thus from the first law:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑄𝑄 = 𝑑𝑑𝑑𝑑 + + 𝑃𝑃 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑇𝑇
• Dividing by 𝑑𝑑𝑑𝑑, we get
𝑑𝑑𝑑𝑑 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑑𝑑𝑑𝑑
= + + 𝑃𝑃
𝑑𝑑𝑑𝑑 𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑇𝑇 𝑑𝑑𝑑𝑑
• This equation is true for any process involving any
temperature change 𝑑𝑑𝑑𝑑 and any volume change
𝑑𝑑𝑑𝑑.
Work and Heat
• Case-A: If 𝑉𝑉 is constant, then 𝑑𝑑𝑑𝑑 = 0, then
𝑑𝑑𝑑𝑑 𝜕𝜕𝜕𝜕
= = 𝐶𝐶𝑉𝑉
𝑑𝑑𝑑𝑑 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑉𝑉
where 𝐶𝐶𝑉𝑉 is the specific heat at constant volume.
• Case-B: If 𝑃𝑃 is constant, then
𝑑𝑑𝑑𝑑 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝑉𝑉
= + + 𝑃𝑃
𝑑𝑑𝑑𝑑 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝑇𝑇 𝑃𝑃
Work and Heat
• But from definition, 𝐶𝐶𝑃𝑃 = 𝑑𝑑𝑑𝑑/𝑑𝑑𝑑𝑑 𝑃𝑃 and we
1 𝜕𝜕𝜕𝜕
got 𝜕𝜕𝜕𝜕/𝜕𝜕𝜕𝜕 𝑃𝑃 = 𝑉𝑉𝑉𝑉, where 𝛽𝛽 = ( ) .
𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕
𝐶𝐶𝑃𝑃 = 𝐶𝐶𝑉𝑉 + + 𝑃𝑃 𝛽𝛽𝛽𝛽
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝐶𝐶𝑃𝑃 −𝐶𝐶𝑉𝑉
Or, = − 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑇𝑇 𝑉𝑉𝑉𝑉
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Problems
• Problem-1: (a) For small changes 𝑑𝑃, 𝑑𝑉 and 𝑑𝑇
of the pressure, 𝑃 , the volume, 𝑉 , and
temperature, 𝑇, in a quasistatic process show
that:
𝑑𝑃 = 𝐵 −𝑑𝑉/𝑉 + 𝛼𝑃 𝑑𝑇
𝜕𝑃
where, 𝐵 = −𝑉 is the isothermal bulk
𝜕𝑉 𝑇
1 𝜕𝑉
modulus, 𝛼𝑃 = is the thermal expansion
𝑉 𝜕𝑇 𝑃
coefficient at constant pressure or the volume
expansivity.
Problems
• Solution-1(a): Every hydrostatic system, has an
equation of state expressing a relation among the
three coordinates (𝑃, 𝑉, 𝑇) , that is valid for all
equilibrium states and during quasi-static processes.
• Hence, we may assume that the pressure is a function
of volume and temperature, i.e.
𝑃 = 𝑃 𝑉, 𝑇
𝜕𝑃 𝜕𝑃
⇒ 𝑑𝑃 = 𝑑𝑉 + 𝑑𝑇
𝜕𝑉 𝑇 𝜕𝑇 𝑉
𝜕𝑃 𝜕𝑇 𝜕𝑉
• Again, from calculus, = −1
𝜕𝑇 𝑉 𝜕𝑉 𝑃 𝜕𝑃 𝑇
Problems
• Solution-1(a)(Contd.): Hence, using
𝜕𝑋/𝜕𝑌 𝑍 = 1/ 𝜕𝑌/𝜕𝑋 𝑍
𝜕𝑃 𝜕𝑉 𝜕𝑃
we get: =−
𝜕𝑇 𝑉 𝜕𝑇 𝑃 𝜕𝑉 𝑇
1 𝜕𝑉 𝜕𝑃
= −𝑉 = 𝐵 𝛼𝑃
𝑉 𝜕𝑇 𝑃 𝜕𝑉 𝑇

• Hence, we get:
𝜕𝑃 𝜕𝑃
𝑑𝑃 = 𝑑𝑉 + 𝑑𝑇
𝜕𝑉 𝑇
𝜕𝑇 𝑉
𝐵
=− 𝑑𝑉 + 𝐵 𝛼𝑃 𝑑𝑇 = 𝐵 −𝑑𝑉/𝑉 + 𝛼𝑃 𝑑𝑇
𝑉
Problems
• Problem-1: (b) When the equation of state is given in
the form 𝑃 = 𝑃(𝑇, 𝑉 ), show that the following
relation holds:
𝑃 𝛼𝑉 = 𝐵 𝛼𝑃
𝜕𝑃
where, 𝐵 = −𝑉 is the isothermal bulk modulus,
𝜕𝑉 𝑇
1 𝜕𝑉
𝛼𝑃 = 𝛽 = is the thermal expansion coefficient at
𝑉 𝜕𝑇 𝑃
constant pressure or the volume expansivity and 𝛼𝑉 =
1 𝜕𝑃
is the thermal coefficient of the pressure at
𝑃 𝜕𝑇 𝑉
constant volume.
Problems
• Solution-1(b): Every hydrostatic system, has an equation of
state expressing a relation among the three coordinates
(𝑃, 𝑉, 𝑇), that is valid for all equilibrium states.
• Hence, we may assume that the pressure is a function of
volume and temperature, i.e.
𝑃 = 𝑃 𝑉, 𝑇
• Hence, from the definitions of 𝛼𝑃 and 𝐵, we get:

𝜕𝑃 𝜕𝑉 𝜕𝑃
𝑃𝛼𝑉 = =−
𝜕𝑇 𝑉
𝜕𝑇 𝑃
𝜕𝑉 𝑇
1 𝜕𝑉 𝜕𝑃
= −𝑉 = 𝐵 𝛼𝑃
𝑉 𝜕𝑇 𝑃
𝜕𝑉 𝑇
Problems
• Problem-2(a): Show that the following relation
exists between the adiabatic compressibility
𝜅𝑎𝑑 = − 1/𝑣 𝜕𝑣/𝜕𝑃 𝑎𝑑 and the isothermal
compressibility 𝜅 𝑇 = − 1/𝑣 𝜕𝑣/𝜕𝑃 𝑇 :
𝑐𝑉
𝜅𝑎𝑑 = 𝜅 𝑇
𝑐𝑃
where, 𝑣 is the molar volume or specific volume
and where 𝑐𝑉 and 𝑐𝑃 are the molar specific heats at
constant volume and constant pressure,
respectively.
Problems
• Strategy for Solution:
• A. The problem involves specific heats at constant
volume and pressure, which are related to the internal
energy 𝑢 = 𝑈/𝑛, i.e. the molar internal energy.
• B. We need to use/derive definitions of 𝑐𝑉 and 𝑐𝑃 , in
terms of rate of change of 𝑢.
• C. For an adiabatic process, use first law with 𝑑𝑞 = 0.
• D. For an isothermal process, use 𝑑𝑇 = 0 and express
𝑇 in terms of the other thermodynamic coordinates.
• E. In each case, get expression of 𝜕𝑃/𝜕𝑣 or 𝜕𝑣/𝜕𝑃 .
Problems
• Solution-2(a): Adiabatic Process:
• We consider hydrostatic system undergoing an
adiabatic quasi-static compression/expansion.
• The molar or specific internal energy 𝑢 may be
thought of as a function of pressure, 𝑃 and
molar volume, 𝑣 i.e. 𝑢 = 𝑢(𝑃, 𝑣)
• For adiabatic process, from the first law:
𝑑𝑢 = 𝑑𝑞𝑖𝑛 + 𝑑𝑤𝑜𝑛 = 𝑑𝑞𝑖𝑛 − 𝑃𝑑𝑣 = −𝑃𝑑𝑣
Problems
• Solution-2(a)(contd.): Thus, 𝑑𝑢 + 𝑃𝑑𝑣 = 0
𝜕𝑢 𝜕𝑢
• From, 𝑑𝑢 = 𝑑𝑢 𝑃, 𝑣 = 𝑑𝑃 + 𝑑𝑣
𝜕𝑃 𝑣 𝜕𝑣 𝑃

𝜕𝑢 𝜕𝑢
• We get, 𝑑𝑃 + + 𝑃 𝑑𝑣 = 0
𝜕𝑃 𝑣 𝜕𝑣 𝑃
• Hence, for an adiabatic process,
𝑑𝑃 𝜕𝑃 𝜕𝑢 𝜕𝑢
= =− +𝑃 /
𝑑𝑣 𝑎𝑑 𝜕𝑣 𝑎𝑑 𝜕𝑣 𝑃 𝜕𝑃 𝑣
Problems
• Solution-2(a)(contd.): Again, for a hydrostatic
system, we may write:
𝑢 = 𝑓 𝑇, 𝑣 = 𝑢(𝑇, 𝑣)
• But, 𝑃, 𝑣, 𝑇 are related, for a hydrostatic system,
by the equation of state.
• Hence, 𝑇 = 𝑇 𝑃, 𝑣 ⇒ 𝑢 = 𝑢(𝑇 𝑃, 𝑣 , 𝑣)
𝜕𝑢 𝜕𝑢 𝜕𝑇
⇒ = by chain rule
𝜕𝑃 𝑣 𝜕𝑇 𝑣 𝜕𝑃 𝑣
• Since at constant 𝑣, 𝑢 may change with respect to
𝑃 through the change of 𝑇.
Problems
• Solution-2(a)(contd.): Again, for a hydrostatic
system, we may write:
𝑢 = 𝑔 𝑇, 𝑃 = 𝑢 𝑇, 𝑃 = 𝑢(𝑇 𝑃, 𝑣 , 𝑃)
𝜕𝑢 𝜕𝑢 𝜕𝑇
⇒ =
𝜕𝑣 𝑃 𝜕𝑇 𝑃 𝜕𝑣 𝑃
• Since at constant 𝑃, 𝑢 may change with respect
to 𝑣 through the change of 𝑇.
𝜕𝑢 𝜕𝑢 𝜕𝑣 𝜕𝑇
⇒ +𝑃 = +𝑃
𝜕𝑣 𝑃 𝜕𝑇 𝑃 𝜕𝑇 𝑃 𝜕𝑣 𝑃
Problems
• Solution-2(a)(contd.): Expressions of 𝑐𝑉 and 𝑐𝑃 :
• From first law, (for a general process)
𝑑𝑢 = 𝑑𝑞𝑖𝑛 + 𝑑𝑤𝑜𝑛 = 𝑑𝑞𝑖𝑛 − 𝑃𝑑𝑣
𝜕𝑢
• A. Hence at constant 𝑣, 𝑐𝑉 = 𝑑𝑞𝑖𝑛 /𝑑𝑇 𝑣 =
𝜕𝑇 𝑣
• B. At constant 𝑃, 𝑐𝑃 = 𝑑𝑞𝑖𝑛 /𝑑𝑇 𝑃 and from first law,
𝜕𝑢 𝑑𝑞𝑖𝑛 𝜕𝑣
= −𝑃
𝜕𝑇 𝑃 𝑑𝑇 𝑃 𝜕𝑇 𝑃
𝜕𝑢 𝜕𝑣
⇒ 𝑐𝑃 = +𝑃
𝜕𝑇 𝑃 𝜕𝑇 𝑃
Problems
• Solution-2(a)(contd.): Express 𝜕𝑃/𝜕𝑣 in terms of 𝑐𝑉 and
𝑐𝑃 :
• We have,
𝜕𝑢 𝜕𝑢 𝜕𝑇 𝜕𝑇
= = 𝑐𝑉
𝜕𝑃 𝑣 𝜕𝑇 𝑣 𝜕𝑃 𝑣 𝜕𝑃 𝑣
𝜕𝑢 𝜕𝑢 𝜕𝑣 𝜕𝑇 𝜕𝑇
+𝑃 = +𝑃 = 𝑐𝑃
𝜕𝑣 𝑃 𝜕𝑇 𝑃 𝜕𝑇 𝑃
𝜕𝑣 𝑃
𝜕𝑣 𝑃
• Hence,
𝜕𝑃 𝜕𝑢 𝜕𝑢 −𝑐𝑃 𝜕𝑇 𝜕𝑇
=− +𝑃 / = /
𝜕𝑣 𝑎𝑑 𝜕𝑣 𝑃 𝜕𝑃 𝑣
𝑐𝑉 𝜕𝑣 𝑃
𝜕𝑃 𝑣
Problems
• Solution-2(a)(contd.): Isothermal Process:
• For an isothermal process, using 𝑇 = 𝑇(𝑃, 𝑣) we get
𝜕𝑇 𝜕𝑇
𝑑𝑇 = 𝑑𝑃 + 𝑑𝑣 = 0
𝜕𝑃 𝑣 𝜕𝑣 𝑃
𝜕𝑇 𝜕𝑇 𝜕𝑣 𝜕𝑇 𝜕𝑃
⇒ =− / =−
𝜕𝑣 𝑃 𝜕𝑃 𝑣 𝜕𝑃 𝑇 𝜕𝑃 𝑣 𝜕𝑣 𝑇
𝜕𝑃 𝑐𝑃 𝜕𝑃
• Hence, = using the above.
𝜕𝑣 𝑎𝑑 𝑐𝑉 𝜕𝑣 𝑇
𝜕𝑣 𝑐𝑉 𝜕𝑣 𝑐𝑉
⇒ = ⇒ 𝜅𝑎𝑑 = 𝜅𝑇
𝜕𝑃 𝑎𝑑
𝑐𝑃 𝜕𝑃 𝑇
𝑐𝑃
(proved)
Problems
• Problem-2(b):
• (i)Derive the compressibility (adiabatic
compressibility), 𝜅𝑎𝑑 , when an ideal gas is quasi-
statically and adiabatically compressed.
• (ii) The velocity of sound is given by 𝑐 =
𝜕𝑃/𝜕𝜌 𝑎𝑑 where, 𝜌 is the density of the gas.
Consider this as a derivative for an adiabatic
change and calculate the sound velocity through
air at 1 atm, 0∘ C, and also the rate of change of
velocity with temperature at 1 atm and 0∘ C.
Problems
• Solution- 2(b): The adiabatic compressibility is
1 𝜕𝑣 1 𝜕𝑉
𝜅𝑎𝑑 = − =−
𝑣 𝜕𝑃 𝑎𝑑 𝑉 𝜕𝑃 𝑎𝑑
• For an ideal gas undergoing adiabatic
compression or expansion, 𝑃𝑉 𝛾 =const.
• Hence, for an adiabatic process of an ideal
𝛾 𝛾−1 𝑑𝑉 −𝑉
gas: 𝑉 𝑑𝑃 + 𝛾𝑉 𝑃 𝑑𝑉 = 0 ⇒ ቚ =
𝑑𝑃 𝑎𝑑 𝛾𝑃
⇒ 𝜅𝑎𝑑 = 1/(𝛾𝑃)
Problems
• Solution- 2(b)(contd.): We can alternatively
𝑐𝑉 −1 𝜕𝑉
derive the same from: 𝜅𝑎𝑑 = 𝜅 𝑇 =
𝑐𝑃 𝛾𝑉 𝜕𝑃 𝑇
𝜕𝑉 𝜕 𝑛𝑅𝑇 𝑛𝑅𝑇 𝑉
• Here, = = − 2 = −
𝜕𝑃 𝑇 𝜕𝑃 𝑃 𝑇 𝑃 𝑃
⇒ 𝜅𝑎𝑑 = 1/(𝛾𝑃)
• The density is: 𝜌 = 𝑚/𝑉 = 𝑛𝑀𝑚𝑜𝑙 /𝑉 = 𝑛𝑀/𝑉,
where 𝑀 is the molecular weight i.e. mass of one
mole of a gas.
𝜌 𝜕𝑉 𝜕𝑉
• Hence, 𝑑𝜌 = − 𝑑𝑉 ⇒ = = −𝑉/𝜌
𝑉 𝜕𝜌 𝜕𝜌 𝑎𝑑
Problems
• Solution- 2(b)(contd.): Hence, (using chin rule)
𝜕𝑃 𝜕𝑃 𝜕𝑉 𝑉 𝜕𝑉
= =− /
𝜕𝜌 𝑎𝑑 𝜕𝑉 𝑎𝑑 𝜕𝜌 𝑎𝑑 𝜌 𝜕𝑃 𝑎𝑑
𝑉 −1
=−
𝜌 𝜅𝑎𝑑 𝑉
• Compression and expansion due to the sound wave
may be considered as quasi-static adiabatic processes.
• The speed of sound:
𝜕𝑃 1 𝛾𝑃 𝛾𝑃𝑉 𝛾𝑅𝑇
𝑐= = = = =
𝜕𝜌 𝑎𝑑 𝜌 𝜅𝑎𝑑 𝜌 𝑛𝑀 𝑀
Problems
• Solution- 2(b)(contd.): The average molecular weight of
air is: 𝑀 = 0.78 × 𝑀𝑁2 + 0.22 × 𝑀𝑂2
• 𝑀 = (0.78 × 28 + 0.22 × 32)g/mol = 28.9g/mol
• We have, 𝛾 = 1.4 for diatomic molecules of air
• Hence, at 𝑇 = 273 K, and using 𝑅 = 8.31 ×
107 erg/(K-mol), the speed of sound at 0∘ C is:
1.4×(8.31×107 )×273
𝑐= cm/sec= 332 m/sec
28.9
• The rate of change of speed with temperature at 0∘ C :
1
𝑑𝑐 1 −2 1
= 𝛾𝑅/𝑀 𝑇 = 𝑐/𝑇 = 0.607m/sec-deg K
𝑑𝑇 2 2
Problems
• Problem-2(c): If 𝑐 is the speed of sound, 𝑐 =
𝜕𝑃/𝜕𝜌 𝑎𝑑 (where 𝑃 is the pressure and 𝜌 is
the density and the partial derivative is for an
adiabatic process) and 𝛾 is the ratio of the
specific heats at constant pressure and at
constant volume, show that the internal energy,
𝑢, and the enthalpy, ℎ, per unit mass of an ideal
gas, with a constant specific heat, may be
expressed by the following expressions:
𝑐2 𝑐2
𝑢= + const. ℎ= + const.
𝛾(𝛾−1) 𝛾−1
Problems
• Solution- 2(c): We have the speed of sound (with 𝑐𝑉
and 𝑐𝑃 as molar specific heats at constant volume
and pressure, respectively):
𝜕𝑃 𝛾𝑅𝑇 𝑐𝑃 − 𝑐𝑉 𝑐𝑉 𝛾𝑇 𝛾 𝛾 − 1 𝑐𝑉 𝑇
𝑐= = = =
𝜕𝜌 𝑎𝑑
𝑀 𝑐𝑉 𝑀 𝑀
𝑐𝑉 𝑇 𝑐2
⇒ 𝑐 2 = 𝛾 𝛾 − 1 𝑐𝑉 𝑇/𝑀 ⇒ =
𝑀 𝛾 𝛾−1
𝑈 𝑈
• The internal energy per unit mass is: 𝑢 = =
𝑚 𝑛𝑀
Problems
• Solution- 2(c)(contd.): But for an ideal gas, molar specific
heat, 𝑐𝑉 is : (Note: 𝐶𝑉 is the heat capacity)
1 𝜕𝑈
𝑐𝑉 = 𝐶𝑉 /𝑛 = ⇒ 𝑈 = 𝑛𝑐𝑉 𝑇 +const.
𝑛 𝜕𝑇 𝑉
• Hence, we have
𝑈 𝑐𝑉 𝑇 𝑐2
𝑢= = +const. = + const.
𝑛𝑀 𝑀 𝛾 𝛾−1
𝐻 𝑃𝑉
• Again, enthalpy per unit mass: ℎ = =𝑢+
𝑚 𝑚
𝑛𝑅𝑇 𝑐𝑃 −𝑐𝑉 𝑇 𝑐𝑉 𝑇 𝑐𝑉 𝑇
⇒ℎ=𝑢+ =𝑢+ = + 𝛾 − 1 +const.
𝑛𝑀 𝑀 𝑀 𝑀
𝑐𝑉 𝑇 𝑐2
⇒ℎ=𝛾 + const. = + const. (proved)
𝑀 𝛾−1
Problems
• Problem-16: Let 𝑑𝑞 be the heat necessary to change
the temperature of a gram of a material by 𝑑𝑇 keeping
the state quantity 𝑥 constant. For the sake of simplicity,
assume that there are only two independent variables,
e.g. the specific volume, 𝑣, and the temperature, 𝑇.
Show that the specific heat 𝑐𝑥 is given by the equation:
𝑑𝑞 𝜕𝑢 𝜕𝑢 𝜕𝑣
𝑐𝑥 = = + +𝑃
𝑑𝑇 𝑥 𝜕𝑇 𝑣 𝜕𝑣 𝑇 𝜕𝑇 𝑥
where, 𝑢 is the internal energy per unit mass and 𝑃 is
the pressure.
Problems
• Solution- 16: From the first law of
thermodynamics, for a hydrostatic system:
𝑑𝑢 = 𝑑𝑞𝑖𝑛 + 𝑑𝑤𝑜𝑛 = 𝑑𝑞 − 𝑃𝑑𝑣 ⇒ 𝑑𝑞 = 𝑑𝑢 + 𝑃𝑑𝑣
• We have, for a hydrostatic system:
𝜕𝑢 𝜕𝑢
𝑢 = 𝑢 𝑇, 𝑣 ⇒ 𝑑𝑢 = 𝑑𝑇 + 𝑑𝑣
𝜕𝑇 𝑣 𝜕𝑣 𝑇
𝜕𝑢 𝜕𝑢
• Hence, 𝑑𝑞 = 𝑑𝑇 + + 𝑃 𝑑𝑣
𝜕𝑇 𝑣 𝜕𝑣 𝑇
• Again, we may consider, 𝑥 = 𝑥(𝑇, 𝑣)
Problems
• Solution-16(contd.):However, for a hydrostatic
system, any of the state variables may be
considered as function of other (two) state
variables. Thus, 𝑣 = 𝑣 𝑇, 𝑥
𝜕𝑣 𝜕𝑣
• Hence, 𝑑𝑣 = 𝑑𝑇 + 𝑑𝑥
𝜕𝑇 𝑥 𝜕𝑥 𝑇
• Hence, at constant 𝑥, where 𝑑𝑥 = 0,
𝑑𝑞 𝜕𝑢 𝜕𝑢 𝜕𝑣
𝑐𝑥 = = + +𝑃
𝑑𝑇 𝑥 𝜕𝑇 𝑣 𝜕𝑣 𝑇 𝜕𝑇 𝑥
(proved)
Problems
• Problem-17: If the specific heat of an ideal gas
for a process where the state variable 𝑥 is kept
constant is 𝑐𝑥 , then show that 𝑃𝑣 𝑓 =const.
𝑐𝑥 −𝑐𝑃
for this process, where 𝑓 is equal to , 𝑃 is
𝑐𝑥 −𝑐𝑉
the pressure, 𝑣 is the specific volume, and 𝑐𝑃
and 𝑐𝑉 are the (constant) specific heats at
constant pressure and constant volume,
respectively.
Problems
• Solution- 17: By definition, the molar specific heat at
constant state variable 𝑥 is:
1 𝑑𝑄
𝑐𝑥 = 𝐶𝑥 /𝑛 ⇒ 𝑐𝑥 =
𝑛 𝑑𝑇 𝑥
• From the first law, for a hydrostatic system
𝑑𝑈 = 𝑑𝑄𝑖𝑛 + 𝑑𝑊𝑜𝑛 = 𝑑𝑄 − 𝑃𝑑𝑉
⇒ 𝑑𝑄 = 𝑑𝑈 + 𝑃𝑑𝑉 ⇒ 𝑑𝑞 = 𝑑𝑢 + 𝑃𝑑𝑣
• But for an ideal gas we have
𝜕𝑢
𝑢 = 𝑢 𝑇 ⇒ 𝑑𝑢 = 𝑑𝑇 = 𝑐𝑉 𝑑𝑇
𝜕𝑇 𝑣
where, 𝑐𝑉 is a constant specific heat of the gas at
constant volume.
Problems
• Solution- 17(contd.): Hence, by definition:
𝜕𝑢
𝑐𝑥 𝑑𝑇 = 𝑑𝑞 ቚ = 𝑑𝑢 + 𝑃𝑑𝑣 𝑥 = 𝑑𝑇 + 𝑃𝑑𝑣
𝑥 𝜕𝑇 𝑣
𝑥
⇒ 𝑐𝑥 𝑑𝑇 = 𝑐𝑉 𝑑𝑇 + 𝑃 𝑑𝑣
where we have dropped the condition for constant 𝑥 on 𝑐𝑉
since it is a constant and on 𝑑𝑇 and 𝑑𝑣 since 𝑇 and 𝑣 are
independent variables.
• For ideal gas, 𝑃𝑣 = 𝑅𝑇 = 𝑐𝑃 − 𝑐𝑉 𝑇
𝑑𝑣
• Hence, 𝑑𝑞 = 𝑐𝑥 𝑑𝑇 = 𝑐𝑉 𝑑𝑇 + 𝑐𝑃 − 𝑐𝑉 𝑇
𝑣
𝑑𝑇 𝑑𝑣
⇒ 𝑐𝑥 − 𝑐𝑉 = 𝑐𝑃 − 𝑐𝑉
𝑇 𝑣
Problems
• Solution- 17(contd.): Again for an ideal gas, 𝑣𝑑𝑃 +
𝑃𝑑𝑣 = 𝑅𝑑𝑇
𝑑𝑃 𝑑𝑣 𝑑𝑇
⇒ + =
𝑃 𝑣 𝑇
• Hence,
𝑑𝑃 𝑑𝑣 𝑑𝑣
𝑐𝑥 − 𝑐𝑉 = 𝑐𝑃 − 𝑐𝑉 − 𝑐𝑥 + 𝑐𝑉 = 𝑐𝑃 − 𝑐𝑥
𝑃 𝑣 𝑣
𝑑𝑃 𝑑𝑣 𝑐𝑥 −𝑐𝑃
• Or, + 𝑓 = 0 where, 𝑓 = .
𝑃 𝑣 𝑐𝑥 −𝑐𝑉
• Integrating we get, ln 𝑃 + 𝑓 ln 𝑣 =const.
⇒ 𝑃𝑣 𝑓 =const. (proved)
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Problems
• Problem-18: Consider a hydrostatic system for which any
two of the pressure, 𝑃𝑃 , the volume, 𝑉𝑉 , and the
temperature, 𝑇𝑇 , may be chosen as the independent
variables to define a thermodynamic equilibrium state.
Denoting the internal energy by 𝑈𝑈 and the enthalpy by
𝐻𝐻 = 𝑈𝑈 + 𝑃𝑃𝑃𝑃 , prove the followings:
• (a) Heat transferred in a quasi-static process:
𝜕𝜕𝜕𝜕 𝜕𝜕𝑉𝑉 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = + 𝑃𝑃 𝑑𝑑𝑑𝑑 + + 𝑃𝑃 𝑑𝑑𝑃𝑃
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝑃𝑃 𝑇𝑇 𝜕𝜕𝑃𝑃 𝑇𝑇
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
= 𝑑𝑑𝑑𝑑 + + 𝑃𝑃 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑃𝑃
Problems
• Strategy for the Solution:
• A. The starting point should be the first law since
it involves 𝑈𝑈 and 𝑄𝑄.
• B. Since in the first relation, the changes are in 𝑇𝑇
and 𝑃𝑃, we must consider 𝑈𝑈 = 𝑈𝑈(𝑇𝑇, 𝑃𝑃)
• C. In parts of the first relation, we have partial
derivative of 𝑉𝑉, with respect to 𝑇𝑇 and 𝑃𝑃. Hence,
we must use 𝑉𝑉 = 𝑉𝑉(𝑇𝑇, 𝑃𝑃) in the expression for
𝑑𝑑𝑑𝑑 in work done.
• D. For the second relation, since the changes are
in 𝑃𝑃 and 𝑉𝑉, we must consider 𝑈𝑈 = 𝑈𝑈(𝑃𝑃, 𝑉𝑉)
Problems
• Solution-18 (a): From the first law we get:
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑄𝑄𝑖𝑖𝑖𝑖 + 𝑑𝑑𝑊𝑊𝑜𝑜𝑜𝑜 = 𝑑𝑑𝑑𝑑 − 𝑃𝑃𝑃𝑃𝑃𝑃
⇒ 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑃𝑃𝑃𝑃𝑃𝑃
• For a hydrostatic system, we may consider
𝑈𝑈 = 𝑈𝑈 𝑇𝑇, 𝑃𝑃 and we get,
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑃𝑃
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝑃𝑃 𝑇𝑇
• Again, we have, 𝑉𝑉 = 𝑉𝑉(𝑇𝑇, 𝑃𝑃) which gives,
𝜕𝜕𝑉𝑉 𝜕𝜕𝑉𝑉
𝑑𝑑𝑉𝑉 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇
Problems
• Solution-18 (a)(contd.):Hence we have, from the first law
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑄𝑄 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑃𝑃 + 𝑃𝑃𝑃𝑃𝑃𝑃
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝑃𝑃 𝑇𝑇
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
⇒ 𝑑𝑑𝑑𝑑 = + 𝑃𝑃 𝑑𝑑𝑑𝑑 + + 𝑃𝑃 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇
which proves the first relation.
• Assuming, 𝑈𝑈 = 𝑈𝑈(𝑃𝑃, 𝑉𝑉), we get using the following in the first law:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑃𝑃 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝑃𝑃 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
⇒ 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑃𝑃 + + 𝑃𝑃 𝑑𝑑𝑑𝑑
𝜕𝜕𝑃𝑃 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃
(proved)
Problems
• Problem-18(b,c): Prove that:
• The specific heat at constant volume:
𝑑𝑑𝑑𝑑 𝜕𝜕𝜕𝜕
𝐶𝐶𝑉𝑉 = =
𝑑𝑑𝑑𝑑 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑉𝑉
• The specific heat at constant pressure:
𝑑𝑑𝑑𝑑 𝜕𝜕𝐻𝐻
𝐶𝐶𝑃𝑃 = =
𝑑𝑑𝑑𝑑 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃
where 𝐻𝐻 = 𝑈𝑈 + 𝑃𝑃𝑃𝑃 is the enthalpy.
• Solution-18(b,c): From the first law, for a hydrostatic
system:
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑄𝑄𝑖𝑖𝑖𝑖 + 𝑑𝑑𝑊𝑊𝑜𝑜𝑜𝑜 = 𝑑𝑑𝑑𝑑 − 𝑃𝑃𝑃𝑃𝑃𝑃
Problems
• Solution-18(b,c)(contd.): Thus, 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑃𝑃𝑃𝑃𝑃𝑃
• Now, the internal energy 𝑈𝑈 may be considered as a
function of any two of the state variables (𝑇𝑇, 𝑉𝑉, 𝑃𝑃).
• Hence, assuming 𝑈𝑈 = 𝑈𝑈(𝑇𝑇, 𝑉𝑉) we get
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑉𝑉
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝑉𝑉 𝑇𝑇
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
⇒ 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + + 𝑃𝑃 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑇𝑇
𝑑𝑑𝑑𝑑 𝜕𝜕𝜕𝜕
• Hence, at constant volume, 𝐶𝐶𝑉𝑉 = =
𝑑𝑑𝑑𝑑 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑉𝑉

(proved)
Problems
• Solution-18(b,c)(contd.): From the definition
of enthalpy and from the first law,
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑃𝑃𝑃𝑃𝑃𝑃 + 𝑉𝑉𝑉𝑉𝑉𝑉 = 𝑑𝑑𝑑𝑑 + 𝑉𝑉𝑉𝑉𝑉𝑉
⇒ 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 − 𝑉𝑉𝑉𝑉𝑉𝑉
• Hence at constant pressure,
𝑑𝑑𝑑𝑑 𝜕𝜕𝜕𝜕
𝐶𝐶𝑃𝑃 = =
𝑑𝑑𝑑𝑑 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃
(proved)
Problems
• Problem-18(d): Prove that:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
(i) = 𝐶𝐶𝑃𝑃 − 𝑃𝑃𝑃𝑃𝑃𝑃 = 𝑉𝑉𝑉𝑉
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝜅𝜅 𝜕𝜕𝜕𝜕
(ii) = 𝐶𝐶𝑉𝑉 = − 𝜅𝜅 𝑃𝑃𝑃𝑃 + 𝐶𝐶𝑃𝑃 /𝛽𝛽
𝜕𝜕𝑃𝑃 𝑉𝑉 𝛽𝛽 𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝑃𝑃
(iii) = 𝑇𝑇 − 𝑃𝑃 = 𝑇𝑇 − 𝑃𝑃
𝜕𝜕𝑉𝑉 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝑇𝑇 𝑉𝑉
• Strategy for the Solution:
• A. In the first relation, we have internal energy 𝑈𝑈,
and heat capacity 𝐶𝐶𝑃𝑃 whose definition involves
𝑑𝑑𝑑𝑑. Hence, we must use the first law.
Problems
• Strategy for the Solution:
• B. In the first relation, we have 𝜕𝜕𝜕𝜕/𝜕𝜕𝜕𝜕 𝑃𝑃 and
𝜕𝜕𝜕𝜕/𝜕𝜕𝑉𝑉 𝑃𝑃 . These suggest that we must use 𝑈𝑈 =
𝑈𝑈(𝑇𝑇, 𝑃𝑃) in the first part and 𝑈𝑈 = 𝑈𝑈(𝑉𝑉, 𝑃𝑃) in the second
part.
• C. The relations involve 𝛽𝛽 = 1/𝑉𝑉 𝜕𝜕𝜕𝜕/𝜕𝜕𝜕𝜕 𝑃𝑃 and
𝜅𝜅 = − 1/𝑉𝑉 𝜕𝜕𝜕𝜕/𝜕𝜕𝜕𝜕 𝑇𝑇 . These suggest, we must use
𝑉𝑉 = 𝑉𝑉(𝑇𝑇, 𝑃𝑃)
• D. The third relation involves separate 𝑇𝑇, which occurs
at the definition of entropy in the form 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑/𝑇𝑇.
Hence, we must use 𝑑𝑑𝑑𝑑 = 𝑇𝑇𝑇𝑇𝑇𝑇 in the first law.
Problems
• Solution-18(d-i): From the first law for a hydrostatic
system: 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 − 𝑃𝑃𝑃𝑃𝑃𝑃, which gives 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑃𝑃𝑃𝑃𝑃𝑃.
• We may consider: 𝑈𝑈 = 𝑈𝑈(𝑇𝑇, 𝑃𝑃) which gives,
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇
• Again, we may consider, 𝑉𝑉 = 𝑉𝑉(𝑇𝑇, 𝑃𝑃), giving
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑 = 𝛽𝛽𝛽𝛽 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇
1 𝜕𝜕𝜕𝜕
where 𝛽𝛽 is the volume expansivity: 𝛽𝛽 =
𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃
Problems
• Solution-18(d-i)(contd.): At constant pressure,
𝑑𝑑𝑑𝑑 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝐶𝐶𝑃𝑃 = = + 𝑃𝑃 = + 𝑃𝑃𝑃𝑃𝑃𝑃
𝑑𝑑𝑑𝑑 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕
⇒ = 𝐶𝐶𝑃𝑃 − 𝑃𝑃𝑃𝑃𝑃𝑃
𝜕𝜕𝜕𝜕 𝑃𝑃
which is the first part of the first relation. (proved)
• Again, we may write, 𝑈𝑈 = 𝑈𝑈(𝑃𝑃, 𝑉𝑉) and 𝑉𝑉 = 𝑉𝑉(𝑇𝑇, 𝑃𝑃) , we get
𝑈𝑈 = 𝑈𝑈(𝑃𝑃, 𝑉𝑉 𝑇𝑇, 𝑃𝑃 ) giving,
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝑉𝑉
=
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝑉𝑉 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃
since, at constant pressure, 𝑃𝑃, 𝑈𝑈 may change with respect to 𝑇𝑇,
through the change of 𝑉𝑉.
Problems
• Solution-18(d-i)(contd.): Hence, we have,
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
= = 𝑉𝑉𝑉𝑉
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
• Hence, = 𝐶𝐶𝑃𝑃 − 𝑃𝑃𝑃𝑃𝑃𝑃 = 𝛽𝛽𝛽𝛽
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃 (proved)
• Solution-18(d-ii): From the first law for a
hydrostatic system: 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 − 𝑃𝑃𝑃𝑃𝑃𝑃
𝑑𝑑𝑑𝑑 𝜕𝜕𝜕𝜕
• 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑃𝑃𝑃𝑃𝑃𝑃 ⇒ 𝐶𝐶𝑉𝑉 = =
𝑑𝑑𝑑𝑑 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑉𝑉
Problems
• Solution-18(d-ii)(contd.): The isothermal
1 𝜕𝜕𝜕𝜕
compressibility 𝜅𝜅 = −
𝑉𝑉 𝜕𝜕𝜕𝜕 𝑇𝑇
1 𝜕𝜕𝜕𝜕
• The volume expansivity is 𝛽𝛽 =
𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃
𝜅𝜅𝐶𝐶𝑉𝑉 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
• Hence, =− /
𝛽𝛽 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃
• But we have from calculus:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
= 1/ and = −1
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃
Problems
• Solution-18(d-ii)(contd.): Hence,
𝜅𝜅𝐶𝐶𝑉𝑉 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
⇒ =− = /
𝛽𝛽 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑉𝑉
𝜅𝜅𝐶𝐶𝑉𝑉 𝜕𝜕𝜕𝜕 𝜕𝜕𝑇𝑇 𝜕𝜕𝜕𝜕
⇒ = =
𝛽𝛽 𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝑃𝑃 𝑉𝑉
𝜕𝜕𝑃𝑃 𝑉𝑉
since, we may consider 𝑈𝑈 = 𝑈𝑈(𝑉𝑉, 𝑇𝑇) and 𝑇𝑇 = 𝑇𝑇(𝑃𝑃, 𝑉𝑉),
giving 𝑈𝑈 = 𝑈𝑈 𝑉𝑉, 𝑇𝑇 𝑃𝑃, 𝑉𝑉 . Hence, a change of 𝑈𝑈 , at
constant 𝑉𝑉, with respect to 𝑃𝑃 occurs only through the
change of 𝑇𝑇, where 𝑇𝑇depends on 𝑃𝑃.
• Now, we may consider, 𝑈𝑈 = 𝑈𝑈(𝑇𝑇, 𝑃𝑃)
Problems
• Solution-18(d-ii)(contd.): Hence,
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑇𝑇 + 𝑑𝑑𝑃𝑃
𝜕𝜕𝑇𝑇 𝑃𝑃 𝜕𝜕𝑃𝑃 𝑇𝑇
• From the first law, for hydrostatic system:
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑃𝑃𝑃𝑃𝑃𝑃
𝑑𝑑𝑑𝑑 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
⇒ 𝐶𝐶𝑃𝑃 = = + 𝑃𝑃 = + 𝑃𝑃𝑃𝑃𝑃𝑃
𝑑𝑑𝑑𝑑 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃
• Again, we may express: 𝑃𝑃 = 𝑃𝑃(𝑇𝑇, 𝑉𝑉) giving
𝜕𝜕𝑃𝑃 𝜕𝜕𝑃𝑃 𝜕𝜕𝜕𝜕 𝜕𝜕𝑉𝑉
𝑑𝑑𝑃𝑃 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑉𝑉 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑/
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝑉𝑉 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝑃𝑃 𝑇𝑇
Problems
• Solution-18(d-ii)(contd.): Hence,
𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 − 𝑑𝑑𝑑𝑑/𝜅𝜅𝜅𝜅
𝜕𝜕𝜕𝜕 𝑉𝑉
• Thus we get,
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑑𝑑𝑑𝑑
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑 −
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉 𝜅𝜅𝜅𝜅
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑑𝑑𝑑𝑑
= + 𝑑𝑑𝑑𝑑 −
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑇𝑇
𝜅𝜅𝜅𝜅
• Hence, from the above line:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝐶𝐶𝑉𝑉 = = +
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉
Problems
• Solution-18(d-ii)(contd.): From the difference of the
expressions for 𝐶𝐶𝑃𝑃 and 𝐶𝐶𝑉𝑉 , we get
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉 = + 𝑃𝑃𝑃𝑃𝑃𝑃 − −
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
= 𝑃𝑃𝑃𝑃𝑃𝑃 −
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉
𝜅𝜅 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜅𝜅
• ⇒ 𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉 = 𝑃𝑃𝑃𝑃𝑃𝑃 −
𝛽𝛽 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉 𝛽𝛽
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝑇𝑇
• But, 𝜅𝜅/𝛽𝛽 = − / =−
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝑉𝑉 𝑃𝑃
Problems
• Solution-18(d-ii)(contd.): Again from calculus:
𝜕𝜕𝜕𝜕 𝜕𝜕𝑇𝑇 𝜕𝜕𝑉𝑉
= −1
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝑉𝑉 𝑃𝑃 𝜕𝜕𝑃𝑃 𝑇𝑇
𝜅𝜅 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜅𝜅
• Hence, 𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉 = 𝑃𝑃𝑃𝑃𝑃𝑃 −
𝛽𝛽 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉 𝛽𝛽
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
= 𝑃𝑃𝑃𝑃𝑃𝑃 +
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕
= 𝑃𝑃𝑃𝑃𝑃𝑃 −
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝜅𝜅
• Hence, = 𝑃𝑃𝑃𝑃𝑃𝑃 − 𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉
𝜕𝜕𝜕𝜕 𝑇𝑇 𝛽𝛽
Problems
• Solution-18(d-ii)(contd.): Hence, we get,
𝜕𝜕𝜕𝜕 𝜅𝜅𝐶𝐶𝑉𝑉 𝜕𝜕𝜕𝜕 𝜅𝜅
= = − 𝑃𝑃𝑃𝑃𝑃𝑃 + 𝐶𝐶𝑃𝑃
𝜕𝜕𝜕𝜕 𝑉𝑉 𝛽𝛽 𝜕𝜕𝜕𝜕 𝑇𝑇 𝛽𝛽
𝜕𝜕𝜕𝜕 𝜅𝜅𝐶𝐶𝑉𝑉 𝜕𝜕𝜕𝜕
⇒ = = − 𝜅𝜅 𝑃𝑃𝑃𝑃 − 𝐶𝐶𝑃𝑃 /𝛽𝛽
𝜕𝜕𝜕𝜕 𝑉𝑉
𝛽𝛽 𝜕𝜕𝜕𝜕 𝑇𝑇
(proved)
Problems
• Solution-18(d-iii): From the first law for a
hydrostatic system: 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 − 𝑃𝑃𝑃𝑃𝑃𝑃
• From the definition of entropy, for a quasi-static
process:
𝑑𝑑𝑑𝑑 = 𝑇𝑇𝑇𝑇𝑇𝑇 ⇒ 𝑑𝑑𝑑𝑑 = 𝑇𝑇𝑇𝑇𝑇𝑇 − 𝑃𝑃𝑃𝑃𝑃𝑃
• Hence, at constant temperature, dividing by 𝑑𝑑𝑑𝑑,
we get
𝜕𝜕𝜕𝜕 𝜕𝜕𝑆𝑆
= 𝑇𝑇 − 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇
Problems
• Solution-18(d-iii)(contd.): From the first law:
1
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑃𝑃𝑃𝑃𝑃𝑃
𝑇𝑇
• From the above, we get:
𝜕𝜕𝜕𝜕 1 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 1 𝜕𝜕𝜕𝜕
= + and =
𝜕𝜕𝜕𝜕 𝑇𝑇 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉
• Now, the condition for 𝑑𝑑𝑑𝑑 to be an exact
differential is:
𝜕𝜕2 𝑆𝑆 𝜕𝜕2 𝑆𝑆
=
𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕 𝜕𝜕𝑉𝑉𝜕𝜕𝑇𝑇
Problems
• Solution-18(d-iii)(contd.): Hence,
𝜕𝜕 1 𝜕𝜕𝜕𝜕 𝜕𝜕 𝑃𝑃 𝜕𝜕 1 𝜕𝜕𝜕𝜕
+ =
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝑉𝑉 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉
• Notice that, the subscripts denote that, the derivatives are
evaluated while keeping that particular variable fixed. But
the result may be a function of that variable.
1 𝜕𝜕𝜕𝜕 1 𝜕𝜕 2 𝑈𝑈 𝜕𝜕 𝑃𝑃 1 𝜕𝜕 2 𝑈𝑈
⇒− 2 + + =
𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝑇𝑇 𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑇𝑇 𝑇𝑇 𝜕𝜕𝑉𝑉𝜕𝜕𝑇𝑇
𝜕𝜕 1/𝑇𝑇
where we have used that = 0 as during the partial
𝜕𝜕𝜕𝜕 𝑇𝑇
derivative with respect to 𝑉𝑉, the other variable 𝑇𝑇 is fixed.
Problems
• Solution-18(d-iii)(contd.): Again, the condition for
𝑑𝑑𝑈𝑈 to be an exact differential is:
𝜕𝜕2 𝑈𝑈 𝜕𝜕2 𝑈𝑈
=
𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕
1 𝜕𝜕𝜕𝜕 𝑃𝑃 1 𝜕𝜕𝜕𝜕
• Hence, − 2 − + =0
𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝑇𝑇 2 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
⇒ = 𝑇𝑇 − 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
• Hence, = 𝑇𝑇 − 𝑃𝑃 = 𝑇𝑇 − 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉

(proved)
Problems
• Problem-18(e): Prove that:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉 = 𝑃𝑃 +
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
=− 𝑃𝑃 +
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇

• Strategy for the Solution:


• A. The obvious starting point would be the first law of
thermodynamics which involves 𝑈𝑈 and using which we
can get 𝐶𝐶𝑉𝑉 and 𝐶𝐶𝑃𝑃 .
Problems
• Strategy for the Solution:
• B. Find expressions for 𝐶𝐶𝑉𝑉 and 𝐶𝐶𝑃𝑃 from the first law.
• C. Partial derivatives of 𝑈𝑈 with respect to 𝑉𝑉 in the first
relation, at constant 𝑇𝑇 , implies we need to take
𝑈𝑈 = 𝑈𝑈 𝑇𝑇, 𝑉𝑉
• D. Partial derivatives of 𝑈𝑈 with respect to 𝑃𝑃 in the
second relation, at constant 𝑇𝑇, implies we need to take
𝑈𝑈 = 𝑈𝑈 𝑇𝑇, 𝑃𝑃
• E. Partial derivative of 𝑉𝑉 with respect to 𝑇𝑇 at constant
𝑃𝑃 implies we need to take 𝑉𝑉 = 𝑉𝑉 𝑇𝑇, 𝑃𝑃 , in both the
relations.
Problems
• Solution-18(e): From the first law we get:
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑄𝑄𝑖𝑖𝑖𝑖 + 𝑑𝑑𝑊𝑊𝑜𝑜𝑜𝑜 = 𝑑𝑑𝑑𝑑 − 𝑃𝑃𝑃𝑃𝑃𝑃 ⇒ 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑃𝑃𝑃𝑃𝑃𝑃
• Hence, at constant 𝑉𝑉:
𝑑𝑑𝑑𝑑 𝜕𝜕𝜕𝜕
𝐶𝐶𝑉𝑉 = =
𝑑𝑑𝑑𝑑 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑉𝑉
• We may take, 𝑈𝑈 = 𝑈𝑈(𝑇𝑇, 𝑉𝑉) giving:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑉𝑉
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝑉𝑉 𝑇𝑇
• Again we may take, 𝑉𝑉 = 𝑉𝑉(𝑇𝑇, 𝑃𝑃) giving:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑃𝑃
𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝑃𝑃 𝑇𝑇
Problems
• Solution-18(e)(contd.): Hence,
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
= + 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇
• From calculus using the chain rule, since 𝑈𝑈 = 𝑈𝑈(𝑇𝑇, 𝑉𝑉(𝑇𝑇, 𝑃𝑃)),
we get:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
=
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝑃𝑃 𝑇𝑇
• Hence,
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = + 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇
Problems
• Solution-18(e)(contd.): Again from the first law:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑃𝑃𝑃𝑃𝑃𝑃 = 𝑑𝑑𝑑𝑑 + 𝑃𝑃 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
= + 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
+ 𝑃𝑃 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
⇒ 𝑑𝑑𝑑𝑑 = + + 𝑃𝑃 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
+ + 𝑃𝑃 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝑇𝑇
Problems
• Solution-18(e)(contd.): Hence, at constant pressure, we get
𝑑𝑑𝑑𝑑 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝐶𝐶𝑃𝑃 = = + + 𝑃𝑃
𝑑𝑑𝑑𝑑 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
⇒ 𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉 = 𝑃𝑃 +
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇
which is the first relation. (proved)
• Again, we may write: 𝑈𝑈 = 𝑈𝑈(𝑇𝑇, 𝑃𝑃) giving
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑃𝑃
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝑃𝑃 𝑇𝑇
• But 𝑃𝑃 may be taken as a function of 𝑇𝑇 and 𝑉𝑉, i.e. 𝑃𝑃 = 𝑃𝑃(𝑇𝑇, 𝑉𝑉):
𝜕𝜕𝑃𝑃 𝜕𝜕𝑃𝑃
𝑑𝑑𝑃𝑃 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑉𝑉
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝑉𝑉 𝑇𝑇
Problems
• Solution-18(e)(contd.): Hence,
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
⇒ 𝑑𝑑𝑑𝑑 = + 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇
• Again from calculus, since, 𝑈𝑈 = 𝑈𝑈(𝑇𝑇, 𝑃𝑃(𝑇𝑇, 𝑉𝑉)), we get (chain rule):
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
=
𝜕𝜕𝑉𝑉 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
⇒ 𝑑𝑑𝑑𝑑 = + 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑇𝑇
Problems
• Solution-18(e)(contd.): Hence, from the first law:
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑃𝑃𝑃𝑃𝑃𝑃
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
= + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕
+ 𝑃𝑃 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑇𝑇
• Hence, at constant 𝑉𝑉, we get:
𝑑𝑑𝑑𝑑 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝐶𝐶𝑉𝑉 = = +
𝑑𝑑𝑑𝑑 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝑉𝑉
Problems
• Solution-18(e)(contd.): Again, we may write 𝑈𝑈 = 𝑈𝑈(𝑇𝑇, 𝑃𝑃)
and 𝑉𝑉 = 𝑉𝑉(𝑇𝑇, 𝑃𝑃) giving
𝜕𝜕𝑈𝑈 𝜕𝜕𝑈𝑈
𝑑𝑑𝑈𝑈 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝑉𝑉 𝜕𝜕𝑉𝑉
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇
• Hence, first law gives,
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑃𝑃𝑃𝑃𝑃𝑃
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
= + 𝑃𝑃 𝑑𝑑𝑑𝑑 + + 𝑃𝑃 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇
Problems
• Solution-18(e)(contd.): Hence, at constant 𝑃𝑃, we get:
𝑑𝑑𝑑𝑑 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝐶𝐶𝑃𝑃 = = + 𝑃𝑃
𝑑𝑑𝑑𝑑 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃
• Hence, subtracting the two expressions:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉 = + 𝑃𝑃 − −
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
=− + 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑃𝑃
• But from calculus:
𝜕𝜕𝜕𝜕 𝜕𝜕𝑇𝑇 𝜕𝜕𝑃𝑃
= −1
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝑃𝑃 𝑉𝑉 𝜕𝜕𝑉𝑉 𝑇𝑇
Problems
• Solution-18(e)(contd.): Hence,
𝜕𝜕𝜕𝜕 𝜕𝜕𝑃𝑃 𝜕𝜕𝑉𝑉
=−
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝑇𝑇 𝑉𝑉
𝜕𝜕𝑃𝑃 𝑇𝑇
• Hence, finally:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉 = − − 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
=− 𝑃𝑃 +
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝑇𝑇
which proves the second relation.
(proved)
Problems
• Problem-18(f): Prove that:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
= + 𝑃𝑃 + 𝑉𝑉 = −𝑇𝑇 + 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑃𝑃
• Strategy for the Solution:
• A. The first relation follows from the definition of enthalpy 𝐻𝐻.
• B. The second relation involves a singular 𝑇𝑇. Hence, we must
use the definition of entropy 𝑆𝑆 in the first law.
• C. The derivative of 𝑉𝑉 with respect to 𝑇𝑇 at constant 𝑃𝑃, implies
we must take 𝑉𝑉 = 𝑉𝑉(𝑇𝑇, 𝑃𝑃).
• D. 𝑉𝑉 = 𝑉𝑉(𝑇𝑇, 𝑃𝑃) implies that we need to take derivatives of 𝑆𝑆
at constant 𝑇𝑇, and constant 𝑃𝑃 i.e. we need to calculate
𝜕𝜕𝜕𝜕/𝜕𝜕𝜕𝜕 𝑇𝑇 and 𝜕𝜕𝜕𝜕/𝜕𝜕𝜕𝜕 𝑇𝑇 and equate 𝜕𝜕 2 𝑆𝑆/𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕 with
𝜕𝜕 2 𝑆𝑆/𝜕𝜕𝑇𝑇𝜕𝜕𝑃𝑃 .
Problems
• Solution-18(f): From the definition of enthalpy:
𝐻𝐻 = 𝑈𝑈 + 𝑃𝑃𝑃𝑃, we get
𝜕𝜕𝜕𝜕 𝜕𝜕𝑈𝑈 𝜕𝜕(𝑃𝑃𝑃𝑃)
= +
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
= + 𝑃𝑃 + 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇
which is the first relation.
• Again, we have from the first law,
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 − 𝑃𝑃𝑃𝑃𝑃𝑃 = 𝑇𝑇𝑇𝑇𝑇𝑇 − 𝑃𝑃𝑃𝑃𝑃𝑃
Problems
• Solution-18(f)(contd.): Hence,
1
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑃𝑃𝑃𝑃𝑃𝑃
𝑇𝑇
• Again, we may consider 𝑉𝑉 = 𝑉𝑉(𝑇𝑇, 𝑃𝑃) giving
𝜕𝜕𝑉𝑉 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑃𝑃
𝜕𝜕𝑇𝑇 𝑃𝑃 𝜕𝜕𝑃𝑃 𝑇𝑇
𝑑𝑑𝑑𝑑 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕
⇒ 𝑑𝑑𝑑𝑑 = + 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝑇𝑇 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑃𝑃 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇
Problems
• Solution-18(f)(contd.): Dividing by 𝑑𝑑𝑑𝑑 and taking
𝑇𝑇 =constant, we get
𝑑𝑑𝑑𝑑 1 𝑑𝑑𝑈𝑈 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑑𝑑𝑇𝑇 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑑𝑑𝑃𝑃
� = � + � + �
𝑑𝑑𝑑𝑑 𝑇𝑇 𝑇𝑇 𝑑𝑑𝑑𝑑 𝑇𝑇 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑃𝑃 𝑑𝑑𝑑𝑑 𝑇𝑇 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝑑𝑑𝑑𝑑 𝑇𝑇
𝜕𝜕𝜕𝜕 1 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕
⇒ = +
𝜕𝜕𝜕𝜕 𝑇𝑇 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
• Hence, + 𝑃𝑃 = 𝑇𝑇
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇
• Again, dividing by 𝑑𝑑𝑇𝑇 and taking 𝑃𝑃 =constant, we get,
𝑑𝑑𝑑𝑑 1 𝑑𝑑𝑑𝑑 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑑𝑑𝑑𝑑
� = � + � +0
𝑑𝑑𝑇𝑇 𝑃𝑃 𝑇𝑇 𝑑𝑑𝑇𝑇 𝑃𝑃 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑃𝑃 𝑑𝑑𝑇𝑇 𝑇𝑇
Problems
• Solution-18(f)(contd.): Hence we get,
𝜕𝜕𝜕𝜕 1 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕
= +
𝜕𝜕𝑇𝑇 𝑃𝑃 𝑇𝑇 𝜕𝜕𝑇𝑇 𝑃𝑃 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑃𝑃
• Now, the condition for 𝑑𝑑𝑑𝑑 to be an exact differential is:
𝜕𝜕 2 𝑆𝑆 𝜕𝜕 2 𝑆𝑆
=
𝜕𝜕𝜕𝜕𝜕𝜕𝑇𝑇 𝜕𝜕𝑇𝑇𝜕𝜕𝜕𝜕
• Hence, we get,
𝜕𝜕 1 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝜕𝜕 1 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕
+ = +
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝑇𝑇 𝑃𝑃 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝑇𝑇 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇
where in the first derivative, with respect to 𝑃𝑃, the other variable 𝑇𝑇 is
held constant, while in the second derivative on the right hand side, 𝑃𝑃
is held constant.
Problems
• Solution-18(f)(contd.): Hence,
1 𝜕𝜕 2 𝑈𝑈 1 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕 2 𝑉𝑉
⇒ + +
𝑇𝑇 𝜕𝜕𝜕𝜕𝜕𝜕𝑇𝑇 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑃𝑃 𝑇𝑇 𝜕𝜕𝑃𝑃𝜕𝜕𝑇𝑇
1 𝜕𝜕𝜕𝜕 1 𝜕𝜕 2 𝑈𝑈 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕 2 𝑉𝑉
=− 2 + − 2 +
𝑇𝑇 𝜕𝜕𝑃𝑃 𝑇𝑇 𝑇𝑇 𝜕𝜕𝑇𝑇𝜕𝜕𝑃𝑃 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝑇𝑇 𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕
• Now, the conditions for 𝑑𝑑𝑑𝑑 and 𝑑𝑑𝑑𝑑 to be exact differentials
are, respectively:
𝜕𝜕 2 𝑈𝑈 𝜕𝜕 2 𝑈𝑈
=
𝜕𝜕𝜕𝜕𝜕𝜕𝑇𝑇 𝜕𝜕𝑇𝑇𝜕𝜕𝜕𝜕
𝜕𝜕 2 𝑉𝑉 𝜕𝜕 2 𝑉𝑉
=
𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕
Problems
• Solution-18(f)(contd.): Hence, we get:
1 𝜕𝜕𝜕𝜕 1 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕
=− 2 − 2
𝑇𝑇 𝜕𝜕𝜕𝜕 𝑃𝑃 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
⇒ + 𝑃𝑃 = −𝑇𝑇
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑃𝑃
• Comparing this with the previously obtained result,
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
+ 𝑃𝑃 = 𝑇𝑇 , we get
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
=−
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑃𝑃
Problems
• Solution-18(f)(contd.): Hence, finally,
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
= + 𝑃𝑃 + 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕
= −𝑇𝑇 + 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑃𝑃
which proves the two parts of the relation.
(proved)
Problems
• Problem-18(g): Prove the following:
𝜕𝜕2 𝑇𝑇 𝜕𝜕𝐶𝐶𝑃𝑃 𝜕𝜕𝜕𝜕 𝜕𝜕𝐶𝐶𝑉𝑉 𝜕𝜕𝜕𝜕
𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉 + − = −1
𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝑉𝑉 𝑃𝑃 𝜕𝜕𝑃𝑃 𝑉𝑉
• Strategy for the Solution:
• A. The expression involves 𝐶𝐶𝑃𝑃 and 𝐶𝐶𝑉𝑉 and their derivatives.
Hence, we must use the first law.
• B. In calculating 𝐶𝐶𝑉𝑉 , take 𝑈𝑈 = 𝑈𝑈 𝑇𝑇, 𝑉𝑉 , while in calculating
𝐶𝐶𝑃𝑃 we may consider 𝑈𝑈 = 𝑈𝑈 𝑇𝑇, 𝑉𝑉(𝑇𝑇, 𝑃𝑃) since we have
derivative of 𝑇𝑇 with respect to 𝑉𝑉 or equivalently, derivative
of 𝑉𝑉 with respect to 𝑇𝑇.
• C. All three terms contain second derivatives; the first term
explicitly and the other two through 𝐶𝐶𝑃𝑃 and 𝐶𝐶𝑉𝑉 .
Problems
• Solution-18(g): Calculation of 𝐶𝐶𝑉𝑉 : From the first law, we
have
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 − 𝑃𝑃𝑃𝑃𝑃𝑃 ⇒ 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑃𝑃𝑃𝑃𝑃𝑃
• We may consider 𝑈𝑈 = 𝑈𝑈(𝑇𝑇, 𝑉𝑉) giving,
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑇𝑇 + 𝑑𝑑𝑉𝑉
𝜕𝜕𝑇𝑇 𝑉𝑉 𝜕𝜕𝑉𝑉 𝑇𝑇
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
• Hence, 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑇𝑇 + + 𝑃𝑃 𝑑𝑑𝑑𝑑
𝜕𝜕𝑇𝑇 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑇𝑇
𝑑𝑑𝑑𝑑 𝜕𝜕𝜕𝜕
⇒ 𝐶𝐶𝑉𝑉 = =
𝑑𝑑𝑑𝑑 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑉𝑉
Problems
• Solution-18(g)(contd.): Calculation of 𝐶𝐶𝑃𝑃 :Again, we may
consider 𝑈𝑈 = 𝑈𝑈(𝑇𝑇, 𝑉𝑉) giving,
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑉𝑉
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝑉𝑉 𝑇𝑇
• Also, we have 𝑉𝑉 = 𝑉𝑉(𝑇𝑇, 𝑃𝑃) giving
𝜕𝜕𝑉𝑉 𝜕𝜕𝑉𝑉
𝑑𝑑𝑉𝑉 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
⇒ 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
= + 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝑇𝑇
Problems
• Solution-18(g)(contd.): From calculus, using the
chain rule, since 𝑈𝑈 = 𝑈𝑈(𝑇𝑇, 𝑉𝑉(𝑇𝑇, 𝑃𝑃)), we get:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝑈𝑈
=
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇
• Hence,
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝑈𝑈
𝑑𝑑𝑑𝑑 = + 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
• Again, 𝑃𝑃𝑃𝑃𝑃𝑃 = 𝑃𝑃 𝑑𝑑𝑑𝑑 + 𝑃𝑃 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇
Problems
• Solution-18(g)(contd.): Hence,
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = + + 𝑃𝑃 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
+ + 𝑃𝑃 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝑇𝑇
𝑑𝑑𝑑𝑑 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
⇒ 𝐶𝐶𝑃𝑃 = = + + 𝑃𝑃
𝑑𝑑𝑑𝑑 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
⇒ 𝐶𝐶𝑃𝑃 = 𝐶𝐶𝑉𝑉 + + 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑇𝑇
Problems
• Solution-18(g): Find the Difference between 𝐶𝐶𝑃𝑃 and 𝐶𝐶𝑉𝑉 :
• Subtracting we get,
𝜕𝜕𝑇𝑇 𝜕𝜕𝜕𝜕
𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉 = + 𝑃𝑃
𝜕𝜕𝑉𝑉 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇
where we have used, 𝜕𝜕𝜕𝜕/𝜕𝜕𝜕𝜕 𝑃𝑃 = 1/ 𝜕𝜕𝜕𝜕/𝜕𝜕𝜕𝜕 𝑃𝑃
• Hence, differentiating with respect to 𝑃𝑃 on both sides, at constant
𝑉𝑉, we get
𝜕𝜕 2 𝑇𝑇 𝜕𝜕𝐶𝐶𝑃𝑃 𝜕𝜕𝜕𝜕 𝜕𝜕𝐶𝐶𝑉𝑉 𝜕𝜕𝜕𝜕
𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉 + −
𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕 𝜕𝜕𝜕𝜕
= +1
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑇𝑇
𝑉𝑉
Problems
𝜕𝜕𝜕𝜕
• Solution-18(g)(contd.): Rearrange terms to get Now, :
𝜕𝜕𝜕𝜕 𝑉𝑉
since, 𝑈𝑈 = 𝑈𝑈(𝑇𝑇, 𝑉𝑉) and 𝐶𝐶𝑉𝑉 = 𝜕𝜕𝜕𝜕/𝜕𝜕𝜕𝜕 𝑉𝑉 , we may consider,
using 𝑇𝑇 = 𝑇𝑇 𝑃𝑃, 𝑉𝑉 :
𝐶𝐶𝑉𝑉 = 𝐶𝐶𝑉𝑉 𝑇𝑇, 𝑉𝑉 = 𝐶𝐶𝑉𝑉 (𝑇𝑇 𝑃𝑃, 𝑉𝑉 , 𝑉𝑉)
𝜕𝜕𝐶𝐶𝑉𝑉 𝜕𝜕𝐶𝐶𝑉𝑉 𝜕𝜕𝜕𝜕
• Hence, =
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝑇𝑇 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕
• Similarly, we may consider = 𝐹𝐹(𝑇𝑇(𝑃𝑃, 𝑉𝑉), 𝑉𝑉) to get,
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕 2 𝑈𝑈 𝜕𝜕𝜕𝜕 𝜕𝜕 2 𝑈𝑈 𝜕𝜕𝜕𝜕
= =
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑇𝑇 𝑉𝑉
𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑉𝑉
since the condition for 𝑑𝑑𝑑𝑑 to be an exact differential is that
the order of differentiation can be interchanged.
Problems
• Solution-18(g)(contd.): Hence, using the definition of 𝐶𝐶𝑉𝑉 , we get
𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝐶𝐶𝑉𝑉 𝜕𝜕𝜕𝜕
=
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉
𝑉𝑉
• Hence,
𝜕𝜕𝐶𝐶𝑉𝑉 𝜕𝜕𝜕𝜕 𝜕𝜕 𝜕𝜕𝜕𝜕
− −
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑇𝑇 𝑉𝑉
𝜕𝜕𝐶𝐶𝑉𝑉 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝐶𝐶𝑉𝑉 𝜕𝜕𝜕𝜕
=− −
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝐶𝐶𝑉𝑉 𝜕𝜕𝜕𝜕 𝜕𝜕𝐶𝐶𝑉𝑉 𝜕𝜕𝜕𝜕
=− +
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝑉𝑉 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝑃𝑃 𝑉𝑉
Problems
• Solution-18(g)(contd.):Again, 𝐶𝐶𝑉𝑉 = 𝐶𝐶𝑉𝑉 𝑉𝑉, 𝑇𝑇 = 𝐶𝐶𝑉𝑉 (𝑉𝑉, 𝑇𝑇 𝑃𝑃, 𝑉𝑉 ) ,
gives,
𝜕𝜕𝐶𝐶𝑉𝑉 𝜕𝜕𝐶𝐶𝑉𝑉
𝑑𝑑𝐶𝐶𝑉𝑉 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝑉𝑉 𝑇𝑇
𝜕𝜕𝑇𝑇 𝑉𝑉
𝜕𝜕𝐶𝐶𝑉𝑉 𝜕𝜕𝜕𝜕 𝜕𝜕𝑇𝑇 𝜕𝜕𝐶𝐶𝑉𝑉
= 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑉𝑉
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝑉𝑉 𝑃𝑃
𝜕𝜕𝑇𝑇 𝑉𝑉
𝜕𝜕𝐶𝐶𝑉𝑉 𝜕𝜕𝐶𝐶𝑉𝑉 𝜕𝜕𝑇𝑇 𝜕𝜕𝐶𝐶𝑉𝑉 𝜕𝜕𝑇𝑇
= + 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑃𝑃
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝑇𝑇 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝑇𝑇 𝑉𝑉 𝜕𝜕𝑃𝑃 𝑉𝑉
• Again we may consider 𝐶𝐶𝑉𝑉 = 𝐶𝐶𝑉𝑉 𝑉𝑉, 𝑃𝑃 which gives
𝜕𝜕𝐶𝐶𝑉𝑉 𝜕𝜕𝐶𝐶𝑉𝑉
𝑑𝑑𝐶𝐶𝑉𝑉 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑃𝑃
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝑃𝑃 𝑉𝑉
Problems
• Solution-18(g)(contd.): Since, 𝑈𝑈 is a state function
and also all of (𝑇𝑇, 𝑉𝑉, 𝑃𝑃), we have 𝑑𝑑𝐶𝐶𝑉𝑉 is an exact
differential.
• Hence, equating the coefficients of the
differentials we get:
𝜕𝜕𝐶𝐶𝑉𝑉 𝜕𝜕𝐶𝐶𝑉𝑉 𝜕𝜕𝐶𝐶𝑉𝑉 𝜕𝜕𝑇𝑇
= +
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝑇𝑇 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝐶𝐶𝑉𝑉 𝜕𝜕𝐶𝐶𝑉𝑉 𝜕𝜕𝑇𝑇
=
𝜕𝜕𝑃𝑃 𝑉𝑉 𝜕𝜕𝑇𝑇 𝑉𝑉 𝜕𝜕𝑃𝑃 𝑉𝑉
Problems
• Solution-18(g)(contd.): Hence, we have
𝜕𝜕𝐶𝐶𝑉𝑉 𝜕𝜕𝜕𝜕 𝜕𝜕 𝜕𝜕𝜕𝜕
− −
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑇𝑇
𝑉𝑉
𝜕𝜕𝐶𝐶𝑉𝑉 𝜕𝜕𝜕𝜕 𝜕𝜕𝐶𝐶𝑉𝑉 𝜕𝜕𝜕𝜕 𝜕𝜕𝐶𝐶𝑉𝑉 𝜕𝜕𝜕𝜕
=− + =−
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑉𝑉
• Hence,
𝜕𝜕 2 𝑇𝑇 𝜕𝜕𝐶𝐶𝑃𝑃 𝜕𝜕𝜕𝜕 𝜕𝜕𝐶𝐶𝑉𝑉 𝜕𝜕𝜕𝜕 𝜕𝜕 𝜕𝜕𝜕𝜕
𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉 + − = +1
𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑇𝑇 𝑉𝑉
𝜕𝜕 2 𝑇𝑇
𝜕𝜕𝐶𝐶𝑃𝑃 𝜕𝜕𝜕𝜕 𝜕𝜕𝐶𝐶𝑉𝑉 𝜕𝜕𝜕𝜕
⇒ 𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉 + − =1
𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑉𝑉
which proves the relation.

(proved)
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Conversion of Work into Heat and
Vice-versa
• Conversion of Work into Heat:
• Consider a thermodynamic system, consisting of an
object (such as a resistor or a piece of stone) immersed
into a medium of large heat capacity, which we call a
reservoir.
• If by mechanical means (by rubbing or friction) or by
electrical means (by passing electricity through it)
some work is done on the object, temperature of the
object tends to rise.
• Any rise of temperature will result in a transfer of
energy, in the form of heat, from the object (system)
to the medium (reservoir) and Vice-verssa.
Conversion of Work into Heat and
Vice-versa
• Conversion of Work into Heat:
• Since the heat capacity of the surrounding
medium (e.g. water) is large, the transfer of heat
to the medium would change the temperature of
the medium only by an infinitesimal amount
(which is negligible).
• Hence, the object would essentially remain at
constant temperature (isothermal work done), so
also the state of the object.
• This may continue for ever (ad infinitum).
Conversion of Work into Heat and
Vice-versa
• Conversion of Work into Heat:
• Hence, we may have a process in which work
is continuously converted into heat without
changing the state of the system.
• In general, work of any kind 𝑊𝑊, may be done
on a system in contact with a reservoir,
causing heat 𝑄𝑄 to leave the system, without
altering the state of the system.
• The system acts merely as an intermediary.
Conversion of Work into Heat and
Vice-versa
• Conversion of Work into Heat:
• From the first law, the work is equal to the
heat, 𝑊𝑊 = 𝑄𝑄; i.e. transformation of work into
heat is accomplished with 100% efficiency.
• Furthermore, this transformation may be
continued indefinitely.
Work → Heat : Efficiency= 100% (possible)
Duration =infinite time (possible)
Conversion of Work into Heat and
Vice-versa
• Conversion of Heat into Work:
• For the conversion of heat into work, we need a
series of processes or a cycle, by which the
conversion may continue indefinitely.
• Also, we need the system to return to its original
state after the end of each cycle.
• In general, a cycle would to convert heat into
work involves:
• A. A heat reservoir at high temperature 𝑇𝑇𝐻𝐻
• B. A heat reservoir at lower temperature 𝑇𝑇𝐿𝐿
Conversion of Work into Heat and
Vice-versa
• Conversion of Heat into Work: Let,
• 𝑄𝑄𝐻𝐻 = heat taken from the high temperature
reservoir by the system
• 𝑄𝑄𝐿𝐿 = heat expelled from the system to the low
temperature reservoir
• 𝑊𝑊 = work done in each cycle
• Then the thermal efficiency of the process is
defined as:
Work output |𝑊𝑊|
𝜂𝜂 = =
Heat input |𝑄𝑄𝐻𝐻 |
Conversion of Work into Heat and
Vice-versa
• Conversion of Heat into Work:
• From the first law: 𝑄𝑄𝐻𝐻 − 𝑄𝑄𝐿𝐿 = 𝑊𝑊 giving
𝑄𝑄𝐿𝐿
𝜂𝜂 = 1 −
𝑄𝑄𝐻𝐻
• Any such cycle (series of processes), where work is
delivered continuously to the surroundings, running the
cycle over and over again, constitutes a heat engine.
• It is a common experience that:
No heat-engine has ever been developed that converts
the heat extracted from a reservoir at a higher
temperature into work without rejecting some heat to
a reservoir at a lower temperature.
The Second Law of Thermodynamics
• The negative statement of impossibility, as observed
in our daily life experience, is the second law of
thermodynamics:
• Kelvin: "It is impossible by means of inanimate material
agency to derive mechanical effect from any portion of
matter by cooling it below the temperature of the
coldest of the surrounding objects."
• Max Planck: "It is impossible to construct an engine
which, working in a complete cycle, will produce no
effect other than, the raising of a weight and the
cooling of a heat reservoir."
The Second Law of Thermodynamics
• Thus the Kelvin-Planck statement of the
second law of thermodynamics is:
It is impossible to construct an engine that,
operating in a cycle, will produce no effect
other than the extraction of heat from a
reservoir and the performance of an equivalent
amount of work.
The Second Law as a New law of
Nature
• Example of Scenarios Violating the Second Law:
• If the second law were not true, it would be possible
to:
• A. Propel a ship across the ocean by extracting heat
from the ocean
• B. Run a power plant by extracting heat from the
surrounding air.
• Notice that, neither of the “impossibilities” violate the
first law of thermodynamics.
• Both the ocean and the surrounding air contain an
enormous store/amount of internal energy, part of
which we may hope to extract in the form of heat.
The Second Law as a New law of
Nature
• The second law, therefore, is not a deduction
from the first law.
• It stands by itself as a separate law of nature.
• Refers to an aspect of nature different from that
described by the first law.
• In addition to the quantity of energy, we must
consider its quality, as indicated by its usefulness.
• The second law refers to the usefulness of
energy.
The Second Law and the Usefulness
Quality of Energy
• As we use energy, its quantity remains
constant, but its quality is degraded.
• Energy degradation is called entropy increase.
• The empirical observation that:
There is always net entropy increase in real
processes
is the second law of thermodynamics.
Energy Degradation ≡ Increase of Entropy
The Second Law and the Usefulness
Quality of Energy
• The first and second laws of thermodynamics allow us
to analyze the limits on the processes that can occur in
nature.
• The first law allows only those processes that conserve
total energy.
• The second law allows only those processes in which
the quality of energy never increases in the system
and surroundings as a whole.
• The property of energy becoming less useful as we use
it is purely a characteristic of the macroscopic realm.
• This behaviour emerges from the microscopic
properties for large number of particles.
The Second Law and the Direction of
Processes in Real World
• What is the criterion for the direction of real processes?
• What do we mean by a real process?
• Obviously, any process that is not ‘‘real’’ cannot occur, at
least it cannot occur in a finite time.
In any real process, there is net degradation of energy.
• According to the second law, processes that result in net
upgrading of energy are not possible.
• The second law thus gives the direction or arrow of time.
• Any process in real macroscopic world that violates the
second law can not run forward in time:
The Second Law ⇒ Arrow of Time
Lack of Proof of the Second Law of
Thermodynamics
• We have not proved and cannot prove the second
law, just as we did not prove the first law.
• These laws are a generalization of our experience,
and no apparent violation of them has ever
survived close scrutiny.
• We must be content to state it as a crystallization
of our experience.
• Both of these laws are fundamental laws of
nature.
The Second Law and the Perpetual
Machine
• Machines that would violate the first law are
called perpetual motion machines of the first
type.
• Machines that violate the second law are called
perpetual motion machines of the second type.
• A perpetual motion machine of the first kind
produces work without the input of energy.
• A perpetual motion machine of the second kind
is a machine which spontaneously converts
thermal energy into mechanical work.
The Second Law and Entropy
• We have defined a new thermodynamic function
that measures energy degradation.
• We call this function the entropy and designate it
by the symbol 𝑆𝑆.
• In terms of entropy our second law becomes:
In any real process, there is net entropy increase.
• The mathematical form of our second law then
becomes:
Δ𝑆𝑆univ = Δ𝑆𝑆sys + Δ𝑆𝑆sur ≥ 0
Entropy as a State Function
• Is entropy a state function?
• There is no reason that the quality of energy
should be less a state function than the quantity
of energy.
• When the system is completely defined, both the
quantity of energy as well as the quality of energy
should be known.
• The state function nature of entropy is critical for
the mathematical definition of this
thermodynamic quantity.
Refrigerators and the Second Law of
Thermodynamics
• Imagine the cycle of a heat engine performed in a
sequence of processes opposite to that of an engine.
• Then some heat is absorbed by the system from a heat
reservoir at a low temperature, a larger amount of heat
is rejected to a heat reservoir at a high temperature.
• Also a net amount of work is done on the system by
the surroundings.
• A machine that performs a cycle in this direction is
called a refrigerator.
• The working substance of a refrigerator is called a
refrigerant.
Refrigerators and the Second Law of
Thermodynamics
• Let, 𝑄𝑄𝐻𝐻 = the amount of heat rejected by the
refrigerant to the high-temperature reservoir;
• 𝑄𝑄𝐿𝐿 = the amount of heat absorbed by the refrigerant
from the low-temperature reservoir; and
• 𝑊𝑊 = the net work done on the refrigerant by the
surroundings.
• Since the refrigerant undergoes a cycle, there is no
change in internal energy.
• From the first law:
𝑄𝑄𝐻𝐻 = 𝑄𝑄𝐿𝐿 + 𝑊𝑊
Refrigerators and the Second Law of
Thermodynamics
• The purpose of a refrigerator is to extract as
much heat |𝑄𝑄𝐿𝐿 | as possible from the low-
temperature reservoir with the expenditure of as
little work |𝑊𝑊| as possible.
• Work is always necessary to transfer heat from a
lower-temperature reservoir to a higher-
temperature reservoir.
• Heat does not flow spontaneously
from low temperature body to a
high temperature body.
Refrigerators and the Second Law of
Thermodynamics
• The negative statement that heat does not
spontaneously flow from low temperature
body to a high temperature body leads us to
the Clausius statement of the second law:
It is impossible to construct a refrigerator that,
operating in a cycle, will produce no effect
other than the transfer of heat from a lower-
temperature reservoir to a higher-temperature
reservoir.
Equivalence of Kelvin-Planck and
Clasius Statements
• Two propositions or statements are said to be
equivalent when:
• The truth of one implies the truth of the second
• And the truth of the second implies the truth of the
first.
𝐾𝐾 = truth of the Kelvin-Planck statement;
−𝐾𝐾 = falsity of the Kelvin-Planck statement;
𝐶𝐶 = truth of the Clausius statement;
−𝐶𝐶 = falsity of the Clausius statement.
• Let, ⊃ means “implies” and ≡ means “equivalence”
Equivalence of Kelvin-Planck and
Clasius Statements
• We wish to prove, 𝐾𝐾 ≡ 𝐶𝐶 which means:
𝐾𝐾 ⊃ 𝐶𝐶 and 𝐶𝐶 ⊃ 𝐾𝐾
• Alternatively, 𝐾𝐾 ≡ 𝐶𝐶 means:
−𝐾𝐾 ⊃ −𝐶𝐶 and −𝐶𝐶 ⊃ −𝐾𝐾
• To prove that −𝐶𝐶 ⊃ −𝐾𝐾, consider a refrigerator, shown
in the left side:

.
Equivalence of Kelvin-Planck and
Clasius Statements
• It requires no work to transfer |𝑄𝑄𝐿𝐿 | units of heat
from a low-temperature reservoir to a high-
temperature reservoir.
• Hence it violates the Clasius statement.
• Suppose that a heat engine (on the right) also
operates between the same two reservoirs.
• Through it the same heat |𝑄𝑄𝐿𝐿 | is delivered to the
low-temperature reservoir.
• This second engine does not
violate any law by itself alone.
Equivalence of Kelvin-Planck and
Clasius Statements
• But the refrigerator and the second engine
together constitute a self-contained machine,
that:
• Takes a (fraction) of heat 𝑄𝑄𝐻𝐻 − |𝑄𝑄𝐿𝐿 | from the
high temperature reservoir and converts all this
heat into work without producing any change in
the low-temperature reservoir.
• Therefore, the refrigerator and engine together
constitute a violation of the Kelvin-Planck
statement.
• Thus, −𝐶𝐶 ⊃ −𝐾𝐾.
Equivalence of Kelvin-Planck and
Clasius Statements
• To prove that −𝐾𝐾 ⊃ −𝐶𝐶, consider an engine,
shown on the left side of the figure below:
• It rejects no heat to the low-temperature
reservoir and that, therefore, violates the Kelvin-
Planck statement.
• Suppose that a refrigerator
on the right also operates
between the same two
reservoirs and uses up all the
work performed by the engine.
Equivalence of Kelvin-Planck and
Clasius Statements
• The refrigerator on the right violates no law.
• But the engine and refrigerator together constitute a
self-contained machine that:
• Transfers heat |𝑄𝑄𝐿𝐿 | from the low-temperature reservoir
to the high-temperature reservoir without producing
any changes elsewhere.
• Therefore, the engine and the
refrigerator together
constitute a violation of the
Clausius statement.
• Thus, −𝐾𝐾 ⊃ −𝐶𝐶
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Heat Engines
• Combustion Engines: The transformation of heat into
work is usually accomplished, in practice, by two
general types of heat engines (combustion engines):
• A. Internal-combustion engine, such as the gasoline
engine and the diesel engine;
• B. External-combustion engine, such as the steam
engine and the Stirling engine.
• In both types of heat engine, a gas or a mixture of
gases is contained in a cylinder, having a movable
piston.
• The cylinder is closed at one end, and a piston is at the
other end.
Heat Engines
• The gas in the confined space is the system.
• The gas undergoes a cycle, thereby causing a
reciprocating piston to impart a rotational motion
to a shaft, which acts against an opposing force.
• It is necessary, in all engines, that:
• A. the gas in the confined space, at some time in
the cycle, be raised to a high temperature and a
high pressure,
• B. the pressure provides the force that performs
external work.
Heat Engines
• Internal Combustion Engines: In the gasoline and diesel
engines, the rapid burning of the fuel and oxygen from
the air takes place within a confined space.
• The confined space is called the combustion chamber.
• Temperature and pressure of the gas is raised inside the
chamber.
• External Combustion Engines: In the steam and Stirling
engines, the increase in temperature and pressure of the
gas is accomplished by high-temperature surroundings.
• The surroundings transfer heat to the system that is
situated inside the chamber.
Heat Engines
• (a) Reciprocating internal combustion engine:
• (b) External combustion engine:

(a) (b)
Heat Engines
• Gasoline Engine: Here, the cycle involves the
performance of six processes.
• Four of the processes in the cycle require
(vertical/horizontal) motion of the piston and are
hence called strokes.
• The system is a mixture of gasoline-vapor and air.
• For a given cycle, the piston moves up and down twice.
• This represents a four-stroke cycle consisting of two
up-strokes and two down-strokes.
• 1. Intake stroke: The mixture moves into the cylinder
due to suction, as the receding piston enlarges the
accessible volume.
Heat Engines
• The outside pressure is greater than the pressure in the
cylinder, so the mixture is pushed into the combustion
chamber.
• This is the energy-input-part of the cycle—energy
enters the system (in the interior of the cylinder) in the
form of potential energy stored in the fuel.
• 2. Compression stroke: The mixture of gasoline-vapor
and air is compressed, until its pressure and
temperature rises considerably.
• This is accomplished by the advancing piston, which
decreases the volume of the combustion chamber.
Heat Engines
• 3. Combustion: Burning of the hot mixture occurs
very rapidly after ignition by an electric spark.
• The resulting combustion products attain a very
high pressure and temperature, but the volume
remains unchanged during this very short time
interval.
• The piston remains essentially motionless during
this very rapid process.
• Since the piston does not move, this process is
not a stroke.
Heat Engines
• First four processes in a gasoline engine:

Intake stroke Compression stroke Power stroke


and combustion
Heat Engines
• 4. Power stroke: The hot combustion products expand
and push the piston away, thus increasing the volume
and decreasing the pressure and temperature.
• The system, acting through the piston, performs work
on the surroundings (crankshaft, transmission, etc.).
• 5. Exhaust: The combustion products at the end of the
power stroke are still at a higher pressure and
temperature than those of the surroundings.
• An exhaust valve allows some gas to escape until the
pressure drops almost to the atmospheric pressure.
• The piston remains essentially motionless during this
process. This is again not a stroke.
Heat Engines
• 6. Exhaust stroke: The piston pushes almost all
the remaining combustion products out of the
cylinder by exerting
a pressure significantly
larger than atmospheric
pressure.
• This is the second up-
stroke of the cycle.
Exhaust Exhaust stroke
Heat Engines
• In the above processes, there are several
phenomena that render an exact mathematical
analysis quite difficult.
• Among these are friction, turbulence, loss of heat
by conduction, and the chemical reaction
between gasoline-vapor and oxygen, etc.
• Otto Cycle: A drastic but useful simplification of
the gasoline engine is provided by neglecting the
troublesome effects of the gasoline engine.
• The Otto cycle is an idealized gasoline engine.
Otto Cycle
• The cycle is named after the German engineer
Nikolaus Otto for his invention in 1876, but
the idea for a four-stroke engine came from
the Frenchman Alphonse Beau de Rochas in
1862.

Alphonse Beau de Rochas Nikolaus August Otto Otto's 1876 four cycle engine
Otto Cycle
• The idealization in Otto cycle is done by setting a set of
conditions:
• 1. The working substance is at all times air, which behaves like
an ideal gas with constant heat capacities.
• 2. All processes are quasi-static.
• 3. There is no friction or turbulence.
• 4. There is no loss of heat through the walls of the combustion
chamber.
• 5. The processes are reversible.
• These assumptions then lead to the idealized air-standard
Otto cycle, which is composed of six simple processes of an
ideal gas.
Otto Cycle
• The six processes of an Otto cycle are shown
below:

1. 5 → 1 Isobaric intake stroke


2. 1 → 2 Adiabatic compression stroke
3. 2 → 3 Isochoric combustion
4. 3 → 4 Adiabatic expansion stroke
5. 4 → 1 Isochoric exhaust
6. 1 → 5 Isobaric exhaust stroke
Otto Cycle
• Process 5 → 1- Intake Stroke : It represents a
quasi-static intake stroke, isobaric at the
atmospheric pressure.
• The volume of the combustion chamber 𝑉𝑉, varies
from zero to 𝑉𝑉1 as the number of moles 𝑛𝑛′ varies
from zero to 𝑛𝑛, according to the equation:
𝑉𝑉1 𝑛𝑛 𝑅𝑅𝑇𝑇1
𝑉𝑉1 = ∫0 𝑑𝑑𝑑𝑑 = ∫0 𝑃𝑃 𝑑𝑑𝑑𝑑 ⇒ 𝑃𝑃0 𝑉𝑉1 = 𝑛𝑛𝑛𝑛𝑇𝑇1
0
where 𝑃𝑃0 is atmospheric pressure and 𝑇𝑇1 is the
temperature of the outside air.
Otto Cycle
• Process 1 → 2- Compression Stroke: It represents a
quasi-static, adiabatic compression stroke.
• There is no friction, and no loss of heat through the
cylinder’s wall.
• The temperature rises from the ambient 𝑇𝑇1 to 𝑇𝑇2 ,
according to the equation:
𝛾𝛾−1 𝛾𝛾−1
𝑇𝑇1 𝑉𝑉1 = 𝑇𝑇2 𝑉𝑉2
where, 𝑉𝑉1 = the larger volume when the piston is at
the bottom of the compression stroke
𝑉𝑉2 = the smaller volume when the piston is at the top.
• The ratio of heat capacities is assumed to be constant.
Otto Cycle
• Process 2 → 3- Combustion: It represents a quasi-
static isochoric increase of temperature and pressure
of 𝑛𝑛 moles of air.
• The increase is imagined to be brought about by an
absorption of heat |𝑄𝑄𝐻𝐻 | from a series of external high-
temperature reservoirs whose temperatures range
from 𝑇𝑇2 to 𝑇𝑇3 .
• If there were only one reservoir at the temperature 𝑇𝑇3 ,
then the flow of heat would not be quasi-static,
because there would be a substantial difference in
temperature between the system and the single
reservoir at 𝑇𝑇3 .
Otto Cycle
• This process is meant to approximate the effect of the
combustion in a gasoline engine when the piston is
essentially motion-less at the top of the stroke.
• Process 3 → 4 - Power Stroke: It represents a quasi-
static adiabatic power stroke, involving a drop in
temperature from 𝑇𝑇3 to 𝑇𝑇4 , according to the equation:
𝛾𝛾−1 𝛾𝛾−1
𝑇𝑇3 𝑉𝑉2 = 𝑇𝑇4 𝑉𝑉1
where 𝑉𝑉4 = 𝑉𝑉1 is larger than 𝑉𝑉2 .
• This process represents the power stroke.
Otto Cycle
• Process 4 → 1 - Exhaust: It represents a quasi-static
isochoric drop in temperature and pressure of 𝑛𝑛 moles of
air. The number of moles 𝑛𝑛′ decreases from 𝑛𝑛 to some
lower value 𝑛𝑛4 .
• The drop is brought about by a rejection of heat |𝑄𝑄𝐿𝐿 | to a
series of low-temperature external reservoirs ranging in
temperature from 𝑇𝑇4 to 𝑇𝑇1 , where 𝑇𝑇1 is the temperature of
the outside air.
• This process is meant to approximate the drop to
atmospheric pressure upon opening the exhaust valve.
• In reality, the temperature does not actually drop to the
temperature of the outside air as it leaves the exhaust port.
Otto Cycle
• Process 1 → 5: Exhaust Stroke: It represents a quasi-
static exhaust stroke, isobaric at atmospheric pressure.
• The volume varies from 𝑉𝑉1 to zero as the number of
moles of exhaust gas varies from 𝑛𝑛4 to zero.
• The temperature remains constant at the value
𝑇𝑇1 = 𝑇𝑇𝑎𝑎𝑎𝑎𝑎𝑎 .
• The two isobaric processes 5 → 1 and 1 → 5 obviously
cancel each other and need not be considered further.
• Of the four remaining processes, only two involve a
flow of heat.
Efficiency of Otto Cycle
• There is an absorption of |𝑄𝑄𝐻𝐻 | units of heat at high
temperatures from 2 → 3, and a rejection of |𝑄𝑄𝐿𝐿 | units of
heat at lower temperatures from 4 → 1.
• Assuming 𝐶𝐶𝑉𝑉 to be constant along the line 2 → 3, we find
for heat entering the system:
𝑇𝑇3
𝑄𝑄𝐻𝐻 = � 𝐶𝐶𝑉𝑉 𝑑𝑑𝑑𝑑 = 𝐶𝐶𝑉𝑉 (𝑇𝑇3 − 𝑇𝑇2 )
𝑇𝑇2
• Similarly, for process 4 → 1, heat leaving the system,
𝑇𝑇1
𝑄𝑄𝐿𝐿 = − � 𝐶𝐶𝑉𝑉 𝑑𝑑𝑑𝑑 = 𝐶𝐶𝑉𝑉 (𝑇𝑇4 − 𝑇𝑇1 )
𝑇𝑇4
𝑄𝑄𝐿𝐿 𝑇𝑇4 −𝑇𝑇1
• The thermal efficiency = 𝜂𝜂 = 1 − =1−
𝑄𝑄𝐻𝐻 𝑇𝑇3 −𝑇𝑇2
Efficiency of Otto Cycle
• The two adiabatic processes during the compression stroke
and power stroke are given by:
𝛾𝛾−1 𝛾𝛾−1
𝑇𝑇1 𝑉𝑉1 = 𝑇𝑇2 𝑉𝑉2
𝛾𝛾−1 𝛾𝛾−1
𝑇𝑇4 𝑉𝑉1 = 𝑇𝑇3 𝑉𝑉2
𝑇𝑇1 𝑇𝑇2 𝑇𝑇4 −𝑇𝑇1 𝑇𝑇3 −𝑇𝑇2
• Hence, = which implies, =
𝑇𝑇4 𝑇𝑇3 𝑇𝑇4 𝑇𝑇3
𝑇𝑇 −𝑇𝑇 𝑇𝑇
which gives, 4 1 = 4
𝑇𝑇3 −𝑇𝑇2 𝑇𝑇3
𝑇𝑇1 𝑇𝑇4 1
• Hence, 𝜂𝜂 = 1 − = 1 − = 1 − = 1 − 𝑟𝑟1−𝛾𝛾
𝑇𝑇2 𝑇𝑇3 𝑉𝑉1 /𝑉𝑉2 𝛾𝛾−1
where 𝑇𝑇1 and 𝑇𝑇2 are the temperatures at the beginning
and end of the compression stroke.
𝑉𝑉1 /𝑉𝑉2 = 𝑟𝑟 is the compression ratio.
Parameters of Internal Combustion
Engines
• Displacement Volume: The displacement volume
of an internal combustion engine is the volume
displaced by the piston as it moves from the
bottom to the top of the cylinder.
• For the gasoline engine, the displacement volume
= 𝑉𝑉1 − 𝑉𝑉2
• Compression Ratio - r: The compression ratio is
the ratio of the maximum and minimum volumes
of the cylinder.
• For the gasoline engine 𝑟𝑟 = 𝑉𝑉1 /𝑉𝑉2
Power Delivered in Otto Cycle
• Example -1 : A six-cylinder gasoline engine has a
displacement volume of 3.00 L operating at 4000
rpm and having a compression ratio of 𝑟𝑟 = 9.50.
The air–fuel mixture enters a cylinder at
atmospheric pressure and an ambient
temperature of 27∘ C. During combustion, the
mixture reaches a temperature of 1350∘ C.
Calculate the power delivered by the engine.
(Take for the gasoline-air mixture: the molar mass
= 28.608 g/mol, specific heat 𝐶𝐶𝑉𝑉 = 0.718kJ/kg.K )
Power Delivered in Otto Cycle
• Solution -1 : The atmospheric pressure 𝑃𝑃0 = 1
atm = 1.013 × 105 Pa ≈ 101.3 kPa.
• The initial temperature 𝑇𝑇1 = 300 K.
𝑉𝑉1
• The compression ratio 𝑟𝑟 = = 9.50
𝑉𝑉2
• The difference of volume is the displacement
volume. For a six-cylinder engine, the total
displacement volume is 3.00 L.
3.00
• Hence, for each cylinder: 𝑉𝑉1 − 𝑉𝑉2 = L = 0.5 L
6
Power Delivered in Otto Cycle
• Hence, 𝑉𝑉1 = 9.5 𝑉𝑉2 ⇒ 𝑉𝑉1 − 𝑉𝑉2 = 8.5 𝑉𝑉2 = 0.5 L
0.5
• ⇒ V2 = L = 0.05882 L= 5.882 × 10−5 𝑚𝑚3 (since 1
8.5 −3
L= 1.0 × 10 𝑚𝑚3 )
• Hence, 𝑉𝑉1 = 5.586 × 10−4 𝑚𝑚3
• The number of moles of air-fuel mixture:
𝑃𝑃1 𝑉𝑉1 1.013 × 105 𝑃𝑃𝑃𝑃 × 5.586 × 10−4 𝑚𝑚3
𝑛𝑛 = =
𝑅𝑅𝑇𝑇1 8.31446 𝐽𝐽/ 𝑚𝑚𝑚𝑚𝑚𝑚. 𝐾𝐾 × 300 𝐾𝐾
= 0.02268585
• Molar mass of gasoline-air mixture = 𝑀𝑀 = 28.608 g/mol
𝑔𝑔
• Mass of air-gasoline = 28.608 × 0.02268585 𝑚𝑚𝑚𝑚𝑚𝑚 =
𝑚𝑚𝑚𝑚𝑚𝑚
6.48997 × 10−4 kg
Power Delivered in Otto Cycle
• For process 1 → 2 i.e. the adiabatic compression,
𝛾𝛾 𝛾𝛾
𝑃𝑃2 𝑉𝑉2 = 𝑃𝑃1 𝑉𝑉1 which gives:
𝑃𝑃2 = 𝑃𝑃1 𝑉𝑉1 /𝑉𝑉2 𝛾𝛾 = 𝑃𝑃1 𝑟𝑟 𝛾𝛾 = 1.013 × 105 𝑃𝑃𝑃𝑃 9.5 1.4
= 2.3682 × 106 𝑃𝑃𝑃𝑃
• Thus the temperature after the compression is:
𝑃𝑃2 𝑉𝑉2 2.3682×106 𝑃𝑃𝑃𝑃 ×5.882×10−5 𝑚𝑚3
𝑇𝑇2 = = = 738.51K
𝑛𝑛𝑛𝑛 0.0226858×8.31446 𝐽𝐽/ 𝑚𝑚𝑚𝑚𝑚𝑚.𝐾𝐾
• The combustion causes the temperature to increase to
𝑇𝑇3 = 1350∘ C= 1623K
𝑛𝑛𝑛𝑛𝑇𝑇3 𝑛𝑛𝑛𝑛𝑇𝑇3
• The pressure after combustion is 𝑃𝑃3 = =
𝑉𝑉3 𝑉𝑉2
Power Delivered in Otto Cycle
0.0226858×8.31446 𝐽𝐽/ 𝑚𝑚𝑚𝑚𝑚𝑚.𝐾𝐾 ×1623𝐾𝐾
• Thus 𝑃𝑃3 =
6 5.882×10−5 𝑚𝑚3
= 5.204 × 10 𝑃𝑃𝑃𝑃
• After the process 3 → 4, an adiabatic expansion,
𝑉𝑉3 𝛾𝛾
the final pressure becomes : 𝑃𝑃4 = 𝑃𝑃3 =
𝛾𝛾 𝑉𝑉4
𝑉𝑉2
𝑃𝑃3 (𝑉𝑉2 = 𝑉𝑉3 , 𝑉𝑉4 = 𝑉𝑉1 )
𝑉𝑉1
⇒ 𝑃𝑃4 = 5.204 × 106 𝑃𝑃𝑃𝑃 × 𝑟𝑟 −𝛾𝛾 = 2.2259 × 105 𝑃𝑃𝑃𝑃
• Thus the final temperature at the end of the cycle
𝑃𝑃4 𝑉𝑉4
: 𝑇𝑇4 = = 659.2 K
𝑛𝑛𝑛𝑛
Power Delivered in Otto Cycle
• Thus heat absorbed during combustion:
𝑄𝑄ℎ = 𝑄𝑄𝑖𝑖𝑖𝑖 = 𝑚𝑚𝐶𝐶𝑉𝑉 (𝑇𝑇3 − 𝑇𝑇2 )
= 6.49 × 10−4 𝑘𝑘𝑘𝑘 × 0.718𝑘𝑘𝑘𝑘𝑘𝑘𝑔𝑔−1 𝐾𝐾 −1 1623 − 659.2 𝐾𝐾
⇒ 𝑄𝑄ℎ = 0.4491 𝑘𝑘𝑘𝑘
• Heat rejected during exhaust:
𝑄𝑄𝑐𝑐 = 𝑄𝑄𝑜𝑜𝑜𝑜𝑜𝑜 = 𝑚𝑚𝐶𝐶𝑉𝑉 (𝑇𝑇4 − 𝑇𝑇1 )
= 6.49 × 10−4 𝑘𝑘𝑘𝑘 × 0.718𝑘𝑘𝑘𝑘𝑘𝑘𝑔𝑔−1 𝐾𝐾 −1 659.2 − 300 𝐾𝐾
⇒ 𝑄𝑄𝑐𝑐 = 0.1673 𝑘𝑘𝑘𝑘
• Work done: 𝑊𝑊𝑛𝑛𝑛𝑛𝑛𝑛 = 𝑄𝑄𝑖𝑖𝑖𝑖 − 𝑄𝑄𝑜𝑜𝑜𝑜𝑜𝑜 = 0.2818 𝑘𝑘𝑘𝑘
𝑊𝑊𝑛𝑛𝑛𝑛𝑛𝑛
• Efficiency of the gasoline engine: 𝜂𝜂 = = 62%
𝑄𝑄𝑖𝑖𝑖𝑖
• Power is delivered in every other revolution of the crankshaft.
1 4000𝑟𝑟𝑟𝑟𝑟𝑟 1𝑚𝑚𝑚𝑚𝑚𝑚
• Power is: 𝑃𝑃 = 6 𝑟𝑟𝑟𝑟𝑟𝑟 0.2818𝑘𝑘𝑘𝑘 = 56.36𝑘𝑘𝑘𝑘
2 𝑚𝑚𝑚𝑚𝑚𝑚 60𝑠𝑠𝑠𝑠𝑠𝑠
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Diesel Engine
• The Diesel engine is a variant of the gasoline engine
where there is no need of a spark for the combustion.
• Only air is admitted in on the intake stroke.
• The air is compressed adiabatically until the
temperature is high enough to ignite oil that is sprayed
into the cylinder after the compression.
• The rate of supply of oil is adjusted so that combustion
takes place approximately isobarically, the piston
moving out during combustion.
• The rest of the cycle, namely, power stroke, exhaust,
and exhaust stroke, is exactly the same as in the
gasoline engine.
Diesel Engine
• Eliminating the usual troublesome effects of a
gasoline engine and making the same simplifying
assumptions as in the Otto cycle, we get a Diesel
cycle.
• It is known as the air-standard Diesel cycle.
• It is named after Rudolf Diesel
who constructed the first
successful diesel engine using
liquid fuel in 1897.

Rudolf Diesel
Diesel Engine
• Spontaneous ignition temperatures (°C) of some hydrocarbons:
Hydrocarbons Ignition Temperature (°C)
Isooctane 447
Benzene 592
Toluene 568
n-Octane 240
n-Decane 232
n-Hexadecane 230
Methanol 470
Ethanol 392

• Diesel fuel is much like decane and hexadecane and burns without a spark.
• Auto fuel is much like isooctane and benzene and, ideally, will not ignite
until the spark goes off.
Diesel Cycle
• The Otto cycle discussed before is an ideal
cycle for internal combustion spark ignition
reciprocating engines.
• In contrast, the Diesel cycle is an ideal cycle
for the operation of internal combustion
compression ignition reciprocating engines.
• The process 2 → 3 in the Otto cycle is
imagined to be horizontal instead of vertical,
and the resulting cycle is the Diesel cycle.
Diesel Cycle
• In the 𝑃𝑃 − 𝑉𝑉 diagram the Otto and Diesel
cycles are shown below:

The Otto Cycle The Diesel Cycle


Diesel Cycle
• The Diesel cycle consists off:
• Process 5 → 1- Intake Stroke
• Process 1 → 2- Adiabatic Compression Stroke
• Process 2 → 3- Isobaric Combustion
• Process 3 → 4- Adiabatic Power Stroke
• Process 4 → 1- Exhaust
• Process 3 → 4- Power Stroke The Diesel Cycle
Diesel Cycle
• Process 5 → 1- Intake Stroke : It represents a quasi-static
intake of air, isobaric at atmospheric pressure.
• Process 1 → 2- Compression Stroke: It represents a quasi-
static, adiabatic compression of air.
• Process 2 → 3- Isobaric Combustion: It represents a quasi-
static absorption of heat thought to be brought about by
absorption from a series of external reservoirs ranging in
temperature from 𝑇𝑇2 to 𝑇𝑇3 .
• This process is meant to approximate the isobaric burning
of the injected oil.
• Process 3 → 4- Power Stroke: It represents a quasi-static
adiabatic expansion of burnt fuel+air, involving a drop in
temperature from 𝑇𝑇3 to 𝑇𝑇4
Diesel Cycle
• Process 4 → 1 - Exhaust: It represents a quasi-static
isochoric drop in temperature and pressure of 𝑛𝑛 moles of
air. The number of moles 𝑛𝑛′ decreases from 𝑛𝑛 to some
lower value 𝑛𝑛4 .
• The drop is thought to be brought about by a rejection of
heat |𝑄𝑄𝐿𝐿 | to a series of low-temperature external reservoirs
ranging in temperature from 𝑇𝑇4 to 𝑇𝑇1 , where 𝑇𝑇1 is the
temperature of the outside air.
• Process 1 → 5: Exhaust Stroke: It represents a quasi-static
exhaust of burnt fuel+air , isobaric at atmospheric pressure.
• The volume varies from 𝑉𝑉1 to zero as the number of moles
of exhaust gas varies from 𝑛𝑛4 to zero.
• The temperature remains constant at the value 𝑇𝑇1 .
Efficiency of Diesel Cycle
• In the intake stroke (5 → 1), the volume of the combustion
chamber varies from zero to 𝑉𝑉1 as the number of moles of
air varies from zero to 𝑛𝑛, according to the equation:
𝑅𝑅𝑇𝑇1
𝑉𝑉 ∝ 𝑛𝑛 ⇒ 𝑉𝑉 = 𝑛𝑛 or, 𝑃𝑃0 𝑉𝑉 = 𝑛𝑛𝑛𝑛𝑇𝑇1
𝑃𝑃0
where 𝑃𝑃0 is atmospheric pressure and 𝑇𝑇1 is the
temperature of the outside air.
• In the compression stroke (1 → 2), the temperature rises
from the ambient 𝑇𝑇1 to 𝑇𝑇2 , according to the equation:
𝛾𝛾−1 𝛾𝛾−1
𝑇𝑇1 𝑉𝑉1 = 𝑇𝑇2 𝑉𝑉2
where, 𝑉𝑉1 = the larger volume when the piston is at the
bottom of the compression stroke
𝑉𝑉2 = the smaller volume when the piston is at the top.
Efficiency of Diesel Cycle
• In the combustion process (2 → 3), the volume increases
from 𝑉𝑉2 to 𝑉𝑉3 while the pressure remains the same
(isobaric combustion).
• Ideal gas law gives:
𝑇𝑇3 𝑉𝑉3
=
𝑇𝑇2 𝑉𝑉2
• Only two processes (2 → 3 and 4 → 1) involve a flow of
heat.
• In the isobaric combustion stroke (2 → 3), assuming 𝐶𝐶𝑃𝑃 to
be constant in the temperature range, we get
𝑇𝑇3
𝑄𝑄𝐻𝐻 = � 𝐶𝐶𝑃𝑃 𝑑𝑑𝑑𝑑 = 𝐶𝐶𝑃𝑃 (𝑇𝑇3 − 𝑇𝑇2 )
𝑇𝑇2
Efficiency of Diesel Cycle
• For process 4 → 1, i.e. isochoric rejection of
heat at exhaust, the heat leaving the system,
𝑇𝑇1
𝑄𝑄𝐿𝐿 = − � 𝐶𝐶𝑉𝑉 𝑑𝑑𝑑𝑑 = 𝐶𝐶𝑉𝑉 (𝑇𝑇4 − 𝑇𝑇1 )
𝑇𝑇4
• Thus, the thermal efficiency of an idealized
Diesel engine is
𝑄𝑄𝐿𝐿 1 (𝑇𝑇4 − 𝑇𝑇1 )
𝜂𝜂 = 1 − =1−
𝑄𝑄𝐻𝐻 𝛾𝛾 (𝑇𝑇3 − 𝑇𝑇2 )
Efficiency of Diesel Cycle
• The two adiabatic processes during the
compression stroke and power stroke are given
by:
𝛾𝛾−1 𝛾𝛾−1
𝑇𝑇1 𝑉𝑉1 = 𝑇𝑇2 𝑉𝑉2
𝛾𝛾−1 𝛾𝛾−1
𝑇𝑇3 𝑉𝑉3 = 𝑇𝑇4 𝑉𝑉1
𝑇𝑇1 𝑇𝑇2 𝑉𝑉2 𝛾𝛾−1 𝑉𝑉2 𝛾𝛾 𝑉𝑉2 𝛾𝛾
• Hence, = = ⇒ 𝑇𝑇1 = 𝑇𝑇4
𝑇𝑇4 𝑇𝑇3 𝑉𝑉3 𝑉𝑉3 𝑉𝑉3

𝑉𝑉3 𝛾𝛾−1 𝑉𝑉3 𝛾𝛾−1 𝑉𝑉2 𝛾𝛾


• Also, 𝑇𝑇4 = 𝑇𝑇3 ⇒ 𝑇𝑇1 = 𝑇𝑇3
𝑉𝑉1 𝑉𝑉1 𝑉𝑉3
Efficiency of Diesel Cycle
𝛾𝛾 1 𝛾𝛾−1 𝑉𝑉2 𝛾𝛾−1 𝑉𝑉2
• Thus, 𝑇𝑇1 = 𝑇𝑇3 𝑉𝑉2 𝑉𝑉1 = 𝑇𝑇3
𝑉𝑉3 𝑉𝑉1 𝑉𝑉3
• Hence, we get,
𝛾𝛾−1
𝑉𝑉3 𝑉𝑉2 𝛾𝛾−1 𝑉𝑉2
1 𝑉𝑉1 𝑇𝑇3 − 𝑇𝑇
𝑉𝑉1 𝑉𝑉3 3
𝜂𝜂 = 1 −
𝛾𝛾 𝑉𝑉2
𝑇𝑇3 − 𝑇𝑇
𝑉𝑉3 3
𝑉𝑉3 𝛾𝛾 𝑉𝑉2 𝛾𝛾
1 𝑉𝑉1 − 𝑉𝑉1
𝜂𝜂 = 1 −
𝛾𝛾 𝑉𝑉3 𝑉𝑉2

𝑉𝑉1 𝑉𝑉1
Note that: In Zemansky Dittmann 𝑟𝑟𝐸𝐸 = 𝑉𝑉1 /𝑉𝑉3 = 𝑉𝑉4 /𝑉𝑉3 is called correctly the
expansion ratio but wrongly the “cutoff ratio”.
Efficiency of Diesel Cycle
• The efficiency of the Diesel cycle depends upon 𝛾𝛾 = 𝐶𝐶𝑃𝑃 /
𝐶𝐶𝑉𝑉 , the ratio of heat capacities.
• This expression may be transformed into:
1 (𝑇𝑇4 − 𝑇𝑇1 )
𝜂𝜂 = 1 −
𝛾𝛾 (𝑇𝑇3 − 𝑇𝑇2 )
𝑉𝑉3 𝛾𝛾
1 𝑉𝑉2 𝑇𝑇1 − 𝑇𝑇1
𝛾𝛾
1 𝑟𝑟𝐶𝐶 − 1 𝑇𝑇1
⇒ 𝜂𝜂 = 1 − =1−
𝑉𝑉
𝛾𝛾 3 𝑇𝑇 − 𝑇𝑇 𝛾𝛾 𝑟𝑟𝐶𝐶 − 1 𝑇𝑇2
𝑉𝑉2 2 2

where 𝑟𝑟𝐶𝐶 = 𝑉𝑉3 /𝑉𝑉2 is called the cutoff ratio , 𝑇𝑇1 and 𝑇𝑇2 are
the temperatures at the beginning and end of the
compression stroke.
Efficiency of Diesel Cycle
• Interestingly, the efficiency of the Diesel cycle does not
depend on the expansion ratio given by 𝑟𝑟𝐸𝐸 = 𝑉𝑉4 /𝑉𝑉3
• Example-1: A typical Diesel engine has a cut-off ratio of
𝑟𝑟𝐶𝐶 = 5 operating between 𝑇𝑇1 = 300 𝐾𝐾 and 𝑇𝑇2 =
990 𝐾𝐾. Find its efficiency.
1 51.4 −1 300
• Solution-1: We have 𝜂𝜂 = 1 − = 54%
1.4 5−1 990
• The thermal efficiencies of actual diesel engines are, of
course, lower, for the same reasons as for the gasoline
engine.
Parameters of Internal Combustion
Diesel Engines
• Displacement Volume: The displacement volume of
an internal combustion Diesel engine is the volume
displaced by the piston as it moves from the bottom
to the top of the cylinder.
• The displacement volume = 𝑉𝑉1 − 𝑉𝑉2
• Compression Ratio - 𝒓𝒓: The compression ratio is the
ratio of the maximum and minimum volumes of the
cylinder: 𝑟𝑟 = 𝑉𝑉1 /𝑉𝑉2
• The compression ratio is important in terms of
design criteria as this is restricted by space
limitations.
Parameters of Internal Combustion
Diesel Engines
• Cut-off Ratio- 𝒓𝒓𝑪𝑪 : The cut-off ratio is the ratio of the
volumes during constant pressure combustion process:
𝑟𝑟𝐶𝐶 = 𝑉𝑉3 /𝑉𝑉2
• The point at which fuel addition/injection is stopped (i.e.
point 3 ) is called the cut-off point, hence the name.
• The cut-off ratio is the ratio of this fraction of the expansion
stroke to the full stroke
• Expansion Ratio - 𝒓𝒓𝑬𝑬 : This is the ratio of the volume after
and before the adiabatic expansion of the (fuel+air) in the
cylinder: 𝑟𝑟𝐸𝐸 = 𝑉𝑉4 /𝑉𝑉3
• The expansion length of the power stroke is actually utilized
for the adiabatic expansion or the significant work output.
Problems
• Problem-1: Show that the efficiency of the Diesel
cycle can be written as:
𝛾𝛾 𝛾𝛾
1 𝑟𝑟𝐶𝐶 − 1 𝑇𝑇1 1 1 𝑟𝑟𝐶𝐶 − 1
𝜂𝜂 = 1 − =1−
𝛾𝛾 𝑟𝑟𝐶𝐶 − 1 𝑇𝑇2 𝛾𝛾 𝑟𝑟 𝛾𝛾−1 𝑟𝑟𝐶𝐶 − 1
where, 𝑟𝑟 = 𝑉𝑉1 /𝑉𝑉2 is the compression ratio,
𝑟𝑟𝐶𝐶 = 𝑉𝑉3 /𝑉𝑉2 is the cut-off ratio
and 𝑇𝑇1 and 𝑇𝑇2 are the temperatures at the
beginning and end of the compression stroke.
𝛾𝛾−1 𝛾𝛾−1
• Solution Hint: Note that: 𝑇𝑇1 𝑉𝑉1 = 𝑇𝑇2 𝑉𝑉2
Note that: The statement in Zemansky Dittmann that”Interestingly, the efficiency
of the Diesel cycle… does not depend on the compression ratio” is WRONG!
Two-Stroke Diesel Engine
• In conventional diesel engines four strokes of the
piston are needed for the completion of a cycle,
and only one of the four is a power stroke.
• Since only air is compressed in the diesel engine,
it is possible to eliminate the exhaust and intake
strokes and thus complete the cycle in two
strokes.
• In the two-stroke-cycle diesel engine, every other
stroke is a power stroke, and thus the power is
doubled.
Two-Stroke Diesel Engine
• Principle of Two-stroke Diesel Engine:
• At the conclusion of the power stroke, when the
cylinder is full of combustion products, the valve
opens, exhaust takes place until the combustion
products are at atmospheric pressure.
• Then, instead of using the piston itself to exhaust the
remaining gases, fresh air is blown into the cylinder,
replacing the combustion products.
• A blower, operated by the engine itself, is used for this
purpose.
• Thus it accomplishes in one simple operation what
formerly required two separate piston strokes.
Advantages of the Diesel Engine
• 1. The Diesel engines are capable of converting
approximately 45-50 percent of the fuel energy to
power.
• 2. Diesel engines have a higher compression ratio
than gasoline models (between 12-16;1
compared to 8-12:1).
• Diesel fuel auto-Ignites and hence does not
require spark plugs or wires.
• Fuel cost is significantly lower.
• Diesel engines produce significantly less carbon
monoxide than gasoline powered ones do.
Advantages of the Diesel Engine
• Diesel fuel is safer because:
• (a) diesel is fundamentally more difficult to catch on
fire or burn than regular gas and will not normally
explode when exposed to a high heat source.
• (b) since diesel fuel does not release gasoline fumes
like regular fuel, the possibility of fuel vapors igniting
and causing a vehicle fire is much lower.
• Diesel engines always run cooler than their gasoline
counterparts. Since the engines are more efficient,
there is less waste heat created during the engine
operation.
Problems
• Problem-12: (a) Show that the internal energy of
a material whose equation of state has the form
𝑃𝑃 = 𝑓𝑓(𝑉𝑉)𝑇𝑇 is independent of the volume. Here
P is the pressure, 𝑇𝑇 is the absolute temperature,
and 𝑓𝑓(𝑉𝑉) is a function of 𝑉𝑉 only.
• (b) The internal energy, 𝑢𝑢, of a unit volume of a
gas is a function of 𝑇𝑇 only and the equation of
1
state for the gas is given by 𝑃𝑃 = 𝑢𝑢(𝑇𝑇) .
3
Determine the functional form of 𝑢𝑢(𝑇𝑇).
Problems
• Solution-12 (a): From the first law and second
law:
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 − 𝑃𝑃𝑃𝑃𝑃𝑃 = 𝑇𝑇𝑇𝑇𝑇𝑇 − 𝑃𝑃𝑃𝑃𝑃𝑃
1
⇒ 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑃𝑃𝑃𝑃𝑃𝑃
𝑇𝑇
• Hence we get,
𝜕𝜕𝜕𝜕 1 𝜕𝜕𝑈𝑈 𝑃𝑃
= +
𝜕𝜕𝜕𝜕 𝑇𝑇 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝑇𝑇
𝜕𝜕𝜕𝜕 1 𝜕𝜕𝑈𝑈
=
𝜕𝜕𝜕𝜕 𝑉𝑉
𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉
Problems
• Solution-12 (a) (contd.): The condition for 𝑑𝑑𝑑𝑑 to be an
exact differential is:
𝜕𝜕2 𝑆𝑆 𝜕𝜕2 𝑆𝑆
=
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝑉𝑉 𝜕𝜕𝑇𝑇
which gives:
𝜕𝜕 1 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕 1 𝜕𝜕𝜕𝜕
+ =
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕
𝑇𝑇 𝑉𝑉 𝑉𝑉 𝑇𝑇
1 𝜕𝜕𝜕𝜕 1 𝜕𝜕 2 𝑈𝑈 𝜕𝜕 𝑃𝑃 1 𝜕𝜕 2 𝑈𝑈
⇒− 2 + + =
𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇
𝑇𝑇 𝜕𝜕𝑇𝑇 𝜕𝜕𝑉𝑉 𝜕𝜕𝜕𝜕 𝑇𝑇 𝑉𝑉 𝑇𝑇 𝜕𝜕𝑉𝑉 𝜕𝜕𝑇𝑇
• Again, the condition for 𝑑𝑑𝑈𝑈 to be an exact differential is:
𝜕𝜕2 𝑈𝑈 𝜕𝜕2 𝑈𝑈
=
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
Problems
• Solution-12 (a) (contd.):
𝜕𝜕𝜕𝜕 2 𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕
• Hence, = 𝑇𝑇 = 𝑇𝑇 − 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉
• Since, 𝑃𝑃 = 𝑓𝑓 𝑉𝑉 𝑇𝑇, we get
𝜕𝜕𝜕𝜕
= 𝑇𝑇𝑇𝑇 𝑉𝑉 − 𝑃𝑃 = 0
𝜕𝜕𝜕𝜕 𝑇𝑇
• Hence, 𝑈𝑈 does not depend on 𝑉𝑉.
Problems
• Solution-12 (b): Consider a gas with volume 𝑉𝑉 at
temperature 𝑇𝑇.
• The internal energy 𝑈𝑈 can be written as
𝑈𝑈 = 𝑢𝑢𝑢𝑢 = 𝑢𝑢 𝑇𝑇 𝑉𝑉
where 𝑢𝑢 is the internal energy of a unit volume of
the gas and 𝑢𝑢 = 𝑢𝑢(𝑇𝑇) is given.
1
• Also, we have 𝑃𝑃 = 𝑇𝑇 , given.
𝑢𝑢
3
• Now we have from the previous problem,
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑇𝑇 𝑑𝑑𝑑𝑑 1
= 𝑇𝑇 − 𝑃𝑃 = − 𝑢𝑢 𝑇𝑇
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉 3 𝑑𝑑𝑑𝑑 3
Problems
• Hence, we get
𝜕𝜕𝜕𝜕 𝑇𝑇 𝑑𝑑𝑑𝑑 1 4 𝑇𝑇 𝑑𝑑𝑑𝑑
= 𝑢𝑢(𝑇𝑇) = − 𝑢𝑢 𝑇𝑇 ⇒ 𝑢𝑢 𝑇𝑇 =
𝜕𝜕𝜕𝜕 𝑇𝑇 3 𝑑𝑑𝑑𝑑 3 3 3 𝑑𝑑𝑑𝑑
4 𝑑𝑑𝑑𝑑
⇒ 𝑑𝑑𝑑𝑑
= ⇒ ln 𝑢𝑢(𝑇𝑇) − 4 ln 𝑇𝑇 =const.
𝑇𝑇 𝑢𝑢
∴ 𝑢𝑢 𝑇𝑇 = (const. ) × 𝑇𝑇 4
• Thus, 𝑢𝑢 is proportional to the fourth power of 𝑇𝑇.
• A gas of light quanta (photons), i.e. the thermal
radiation field is such a gas.
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Problems
• Problem-12: (a) Show that the internal energy of
a material whose equation of state has the form
𝑃𝑃 = 𝑓𝑓(𝑉𝑉)𝑇𝑇 is independent of the volume. Here
P is the pressure, 𝑇𝑇 is the absolute temperature,
and 𝑓𝑓(𝑉𝑉) is a function of 𝑉𝑉 only.
• (b) The internal energy, 𝑢𝑢, of a unit volume of a
gas is a function of 𝑇𝑇 only and the equation of
1
state for the gas is given by 𝑃𝑃 = 𝑢𝑢(𝑇𝑇) .
3
Determine the functional form of 𝑢𝑢(𝑇𝑇).
Problems
• Solution-12 (a): From the first law and second
law:
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 − 𝑃𝑃𝑃𝑃𝑃𝑃 = 𝑇𝑇𝑇𝑇𝑇𝑇 − 𝑃𝑃𝑃𝑃𝑃𝑃
1
⇒ 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑃𝑃𝑃𝑃𝑃𝑃
𝑇𝑇
• Hence we get,
𝜕𝜕𝜕𝜕 1 𝜕𝜕𝑈𝑈 𝑃𝑃
= +
𝜕𝜕𝜕𝜕 𝑇𝑇 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝑇𝑇
𝜕𝜕𝜕𝜕 1 𝜕𝜕𝑈𝑈
=
𝜕𝜕𝜕𝜕 𝑉𝑉
𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉
Problems
• Solution-12 (a) (contd.): The condition for 𝑑𝑑𝑑𝑑 to be an
exact differential is:
𝜕𝜕2 𝑆𝑆 𝜕𝜕2 𝑆𝑆
=
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝑉𝑉 𝜕𝜕𝑇𝑇
which gives:
𝜕𝜕 1 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕 1 𝜕𝜕𝜕𝜕
+ =
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕
𝑇𝑇 𝑉𝑉 𝑉𝑉 𝑇𝑇
1 𝜕𝜕𝜕𝜕 1 𝜕𝜕 2 𝑈𝑈 𝜕𝜕 𝑃𝑃 1 𝜕𝜕 2 𝑈𝑈
⇒− 2 + + =
𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇
𝑇𝑇 𝜕𝜕𝑇𝑇 𝜕𝜕𝑉𝑉 𝜕𝜕𝜕𝜕 𝑇𝑇 𝑉𝑉 𝑇𝑇 𝜕𝜕𝑉𝑉 𝜕𝜕𝑇𝑇
• Again, the condition for 𝑑𝑑𝑈𝑈 to be an exact differential is:
𝜕𝜕2 𝑈𝑈 𝜕𝜕2 𝑈𝑈
=
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
Problems
• Solution-12 (a) (contd.):
𝜕𝜕𝜕𝜕 2 𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕
• Hence, = 𝑇𝑇 = 𝑇𝑇 − 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉
• Since, 𝑃𝑃 = 𝑓𝑓 𝑉𝑉 𝑇𝑇, we get
𝜕𝜕𝜕𝜕
= 𝑇𝑇𝑇𝑇 𝑉𝑉 − 𝑃𝑃 = 0
𝜕𝜕𝜕𝜕 𝑇𝑇
• Hence, 𝑈𝑈 does not depend on 𝑉𝑉.
Problems
• Solution-12 (b): Consider a gas with volume 𝑉𝑉 at
temperature 𝑇𝑇.
• The internal energy 𝑈𝑈 can be written as
𝑈𝑈 = 𝑢𝑢𝑢𝑢 = 𝑢𝑢 𝑇𝑇 𝑉𝑉
where 𝑢𝑢 is the internal energy of a unit volume of
the gas and 𝑢𝑢 = 𝑢𝑢(𝑇𝑇) is given.
1
• Also, we have 𝑃𝑃 = 𝑇𝑇 , given.
𝑢𝑢
3
• Now we have from the previous problem,
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑇𝑇 𝑑𝑑𝑑𝑑 1
= 𝑇𝑇 − 𝑃𝑃 = − 𝑢𝑢 𝑇𝑇
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉 3 𝑑𝑑𝑑𝑑 3
Problems
• Hence, we get
𝜕𝜕𝜕𝜕 𝑇𝑇 𝑑𝑑𝑑𝑑 1 4 𝑇𝑇 𝑑𝑑𝑑𝑑
= 𝑢𝑢(𝑇𝑇) = − 𝑢𝑢 𝑇𝑇 ⇒ 𝑢𝑢 𝑇𝑇 =
𝜕𝜕𝜕𝜕 𝑇𝑇 3 𝑑𝑑𝑑𝑑 3 3 3 𝑑𝑑𝑑𝑑
4 𝑑𝑑𝑑𝑑
⇒ 𝑑𝑑𝑑𝑑
= ⇒ ln 𝑢𝑢(𝑇𝑇) − 4 ln 𝑇𝑇 =const.
𝑇𝑇 𝑢𝑢
∴ 𝑢𝑢 𝑇𝑇 = (const. ) × 𝑇𝑇 4
• Thus, 𝑢𝑢 is proportional to the fourth power of 𝑇𝑇.
• A gas of light quanta (photons), i.e. the thermal
radiation field is such a gas.
Joule-Brayton Cycle
• Historical Asides: In 1872, an American engineer,
George Bailey Brayton advanced the study of heat
engines by patenting a constant pressure internal
combustion engine, initially using vaporized gas but
later using liquid fuels such as kerosene.
Pressure
Joule-Brayton Cycle
• The original Brayton engines used a piston
compressor and piston expander, but today's
modern gas turbine engines and air-breathing jet
engines are also a constant-pressure heat engines,
therefore we describe their thermodynamics by the
Brayton or Joule-Brayton cycle.
Joule-Brayton Cycle
• Although the cycle is usually run as an open
system, it is conventionally assumed for the
purposes of thermodynamic analysis that the
exhaust gases are reused in the intake, enabling
analysis as a closed system.
• In general, the Brayton cycle describes the
workings of a constant-pressure heat engine.
• Two major applications of the gas turbine engines
are air-craft propulsion and electric power
generation.
Joule-Brayton Cycle
• The main processes of the Brayton cycle are:
• 1 → 2 : Adiabatic, quasi-static (or reversible)
compression in the inlet and compressor.
• 2 → 3 : Constant pressure fuel combustion
(idealized as constant pressure heat addition)
• 3 → 4 : Adiabatic, quasi-static Pressure

(or reversible) expansion in the


turbine and exhaust nozzle.
• 4 → 1 : Cool the air intake at
constant pressure.
Joule-Brayton Cycle
• The work done in the Joule-Brayton cycle is
utilized in two processes:
• A. we take some work out and use it to drive
the compressor.
• B. We use the remaining work to accelerate
fluid for jet propulsion, or to turn a generator
for electrical power generation.
Problems
• Example-2: Brayton Cycle or Joule Cycle : The
Brayton or Joule cycle represents the
operation of a gas-turbine engine. It consists
of two adiabats and two isobars as shown
below. Pressure

Find the efficiency


of the Brayton-Joule
cycle.
Problems
• Solution-2: Heat is absorbed in the process 2 → 3
and is rejected to the atmosphere in process
4 → 1.
• Along the adiabats, there is no heat transfer.
• Work done by engine along isobar 2 → 3 is :
𝑊𝑊2→3 = 𝑃𝑃2 (𝑉𝑉3 − 𝑉𝑉2 )
• Work done on the system (gas)
along isobar 4 → 1 is:
𝑊𝑊4→1 = −𝑃𝑃1 (𝑉𝑉4 − 𝑉𝑉1 )
Problems
• Solution-2 (contd.): Along the aidbats, from first law,
Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑄𝑄𝑖𝑖𝑖𝑖 + 𝑊𝑊𝑜𝑜𝑜𝑜 = 𝑊𝑊𝑜𝑜𝑜𝑜 (since 𝑄𝑄 = 0 along the
adiabats).
• Hence, work done by the system along the adiabats
1 → 2 and 3 → 4 are:
𝑊𝑊1→2 + 𝑊𝑊3→4 = Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,1→2 + Δ𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖,3→4
= −𝐶𝐶𝑉𝑉 𝑇𝑇2 − 𝑇𝑇1 + 𝐶𝐶𝑉𝑉 (𝑇𝑇3 − 𝑇𝑇4 )
• Total work done by the system is: 𝑊𝑊𝑡𝑡𝑡𝑡𝑡𝑡 =
𝐶𝐶𝑉𝑉 𝑇𝑇3 − 𝑇𝑇4 − (𝑇𝑇2 − 𝑇𝑇1 ) + 𝑃𝑃2 (𝑉𝑉3 − 𝑉𝑉2 ) − 𝑃𝑃1 (𝑉𝑉4 − 𝑉𝑉1 )
• Use the ideal gas law: 𝑃𝑃𝑖𝑖 𝑉𝑉𝑖𝑖 = 𝑛𝑛𝑛𝑛𝑇𝑇𝑖𝑖 , 𝑖𝑖 = 1,2,3,4 with
𝑃𝑃2 = 𝑃𝑃3 and 𝑃𝑃4 = 𝑃𝑃1
Problems
• Solution-2 (contd.): Hence,
𝑊𝑊𝑡𝑡𝑡𝑡𝑡𝑡 = 𝑛𝑛𝑛𝑛 𝑇𝑇3 − 𝑇𝑇2 − 𝑇𝑇4 − 𝑇𝑇1
+ 𝐶𝐶𝑉𝑉 𝑇𝑇3 − 𝑇𝑇4 − (𝑇𝑇2 − 𝑇𝑇1 )
⇒ 𝑊𝑊𝑡𝑡𝑡𝑡𝑡𝑡 = (𝐶𝐶𝑉𝑉 + 𝑛𝑛𝑛𝑛) 𝑇𝑇3 − 𝑇𝑇2 − 𝑇𝑇4 − 𝑇𝑇1
• Hence, 𝑊𝑊𝑡𝑡𝑡𝑡𝑡𝑡 = 𝐶𝐶𝑃𝑃 𝑇𝑇3 − 𝑇𝑇2 − 𝑇𝑇4 − 𝑇𝑇1
• Again in the process 2 → 3 heat absorbed is:
𝑄𝑄ℎ = 𝐶𝐶𝑃𝑃 (𝑇𝑇3 − 𝑇𝑇2 )
• For the adiabatic processes: 𝑃𝑃𝑉𝑉 𝛾𝛾 = 𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐. ⇒
𝑃𝑃1−𝛾𝛾 𝑇𝑇 𝛾𝛾 = 𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐. ⇒ 𝑇𝑇𝑃𝑃(1−𝛾𝛾)/𝛾𝛾 = 𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐𝑐.
Problems
• Solution-2 (contd.): Hence,
𝑇𝑇1 𝑃𝑃1 (𝛾𝛾−1)/𝛾𝛾
=
𝑇𝑇2 𝑃𝑃2
𝑇𝑇4 𝑃𝑃4 (𝛾𝛾−1)/𝛾𝛾 𝑃𝑃 (𝛾𝛾−1)/𝛾𝛾
1
= =
𝑇𝑇3 𝑃𝑃3 𝑃𝑃2
• Hence,
𝑃𝑃1 (𝛾𝛾−1)/𝛾𝛾
𝑊𝑊𝑡𝑡𝑡𝑡𝑡𝑡 = 𝐶𝐶𝑃𝑃 𝑇𝑇3 − 𝑇𝑇2 − 𝑇𝑇3 − 𝑇𝑇2
𝑃𝑃2
𝑊𝑊 𝑃𝑃 (𝛾𝛾−1)/𝛾𝛾
1
• Hence the efficiency is: 𝜂𝜂 = =1−
𝑄𝑄ℎ 𝑃𝑃2
Stirling Cycle
• Historical Aside: In 1816, due to explosions of
steam engines and loss of life, a minister of the
Church of Scotland, named Robert Stirling,
designed and patented a hot-air engine that
could convert some of the energy liberated by a
burning fuel into work.
• It is a regenerative heat engine, where some heat
is stored in an internal reservoir (regenerator)
and taken back in another part of the cycle.
Stirling Cycle
• The schematic diagram of the Stirling cycle is
given below:

|𝑄𝑄𝐻𝐻 |

𝑇𝑇𝐻𝐻
𝑇𝑇𝐿𝐿

|𝑄𝑄𝐶𝐶 |
Stirling Cycle
• The cycle consists of four processes, two
isotherms and two isochoric processes:
• 3 → 4: Isothermal expansion: heat addition from
external source.
• 4 → 1 : Constant volume heat transfer: heat
transfer from the gas to the regenerator.
• 1 → 2: Isothermal compression: heat rejection to
the external sink.
• 2 → 3: Constant volume heat transfer: internal
heat transfer from the regenerator to the gas.
Stirling Cycle

3 → 4 𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶 𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆 4 → 1 𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶 𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉

1 → 2 𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶 𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆
2 → 3 𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶𝐶 𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉𝑉 1 → 2 𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸𝐸 𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆𝑆
Problems
• Problem-3: Stirling Cycle: The Stirling cycle
consists of two isotherms and two isochores as
shown in the figure below.
• Find the efficiency
of the Stirling cycle.
• Solution-3:
• Heat is absorbed at the
high temperature 𝑇𝑇𝐻𝐻 and
rejected at the low
temperature 𝑇𝑇𝐿𝐿 .
Problems
• Solution-3 (contd.): Heat rejected to the
regenerator during process 4 → 1.
• The same amount of heat is absorbed from
the regenerator during the process 2 → 3.
• Hence, we need not consider these flows of
heats any more.
• Heat supplied is equal to the work done
during isothermal expansion (as 𝑇𝑇 is constant,
so is 𝑈𝑈 = 𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 for an ideal gas).
Problems
• Solution-3 (contd.): Similarly, heat rejected to air
during isothermal compression is equal to the work
done on the system (gas).
𝑉𝑉4 𝑉𝑉4
• Hence, 𝑄𝑄𝐻𝐻 = 𝑊𝑊𝑏𝑏𝑏𝑏 = + ∫𝑉𝑉 𝑃𝑃𝑃𝑃𝑃𝑃 = 𝑛𝑛𝑛𝑛𝑇𝑇𝐻𝐻 ∫𝑉𝑉 𝑑𝑑𝑑𝑑/𝑉𝑉
3 3
𝑉𝑉4
⇒ 𝑄𝑄𝐻𝐻 = 𝑛𝑛𝑛𝑛𝑇𝑇𝐻𝐻 ln
𝑉𝑉3
𝑉𝑉2
• Similarly, 𝑄𝑄𝐿𝐿 = −𝑛𝑛𝑛𝑛𝑇𝑇𝐿𝐿 ln
𝑉𝑉1
• From the first law,
Net work done=heat supplied –heat rejected
Problems
• Solution-3 (contd.): Hence, efficiency of the
Stirling cycle is:
Net work done 𝑄𝑄𝐿𝐿 𝑇𝑇𝐿𝐿
𝜂𝜂 = =1− =1−
Heat supplied 𝑄𝑄𝐻𝐻 𝑇𝑇𝐻𝐻
𝑇𝑇1
𝜂𝜂 = 1 −
𝑇𝑇3
Mathematical Theorems
• Differentials: Consider an inclined plane in the
𝑥𝑥 − 𝑦𝑦 − 𝑧𝑧 coordinate space:
• We can think the plane as a
meadow with
• x-axis pointed east
• y-axis pointed north
• z-coordinate is the
elevation above the sea level.
• Elevation is a state function
of the east–north position in the field, i.e.
𝑧𝑧 = 𝑧𝑧(𝑥𝑥, 𝑦𝑦)
Mathematical Theorems
• We seek change of elevation (𝑧𝑧) as we go from point 𝑖𝑖 to
point 𝑓𝑓 in the meadow.
• This can be calculated by first traveling east, a distance Δ𝑥𝑥,
ending up at point 𝑔𝑔.
• The easterly slope of the meadow (a plane) is: Δ𝑧𝑧/Δ𝑥𝑥 𝑦𝑦
• Change in elevation upon arrival at point 𝑔𝑔 is:
(Δ𝑧𝑧)1 = Δ𝑧𝑧/Δ𝑥𝑥 𝑦𝑦 Δ𝑥𝑥
• We complete our journey to point 𝑓𝑓 by traveling in the
northerly, by a distance Δy.
• The northerly slope of the meadow (a plane) is: Δ𝑧𝑧/Δ𝑦𝑦 𝑥𝑥
• Change in elevation in the second part of the trip is:
(Δ𝑧𝑧)2 = Δ𝑧𝑧/Δ𝑦𝑦 𝑥𝑥 Δy
Mathematical Theorems
• The change in elevation for the total trip is the
sum of that in the two parts:
Δ𝑧𝑧 = Δ𝑧𝑧 𝑡𝑡𝑡𝑡𝑡𝑡 = (Δ𝑧𝑧)1 + (Δ𝑧𝑧)2
Δ𝑧𝑧 Δ𝑧𝑧
⇒ Δ𝑧𝑧 = Δ𝑥𝑥 + Δy
Δ𝑥𝑥 𝑦𝑦 Δ𝑦𝑦 𝑥𝑥
• Since the meadow is planar, we could have taken
the journey in the alternate order i.e. going
northerly first and then easterly and would have
got the same result.
Mathematical Theorems
• What happens if the meadow is not planar?
• At a given point, we can approximate a nonplanar
meadow with the plane that is tangent to the meadow
at the point.
• If we stay very close to the original point (i.e. 𝑓𝑓 is not
far from 𝑖𝑖 ), the tangent plane is a very good
approximation to the meadow.
• In the limit of infinitesimal displacement from the
point, it is a perfect approximation and we get:
Δ𝑧𝑧 → 𝑑𝑑𝑑𝑑, Δ𝑥𝑥 → 𝑑𝑑𝑑𝑑, Δ𝑦𝑦 → 𝑦𝑦
Δ𝑧𝑧 𝜕𝜕𝜕𝜕 Δ𝑧𝑧 𝜕𝜕𝜕𝜕
𝐿𝐿𝐿𝐿𝑚𝑚Δ𝑥𝑥→0 = , 𝐿𝐿𝐿𝐿𝑚𝑚Δ𝑦𝑦→0 =
Δ𝑥𝑥 𝑦𝑦 𝜕𝜕𝜕𝜕 𝑦𝑦 Δ𝑦𝑦 𝑥𝑥 𝜕𝜕𝑦𝑦 𝑥𝑥
Mathematical Theorems
• Hence we get,
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑦𝑦
𝜕𝜕𝜕𝜕 𝑥𝑥
• 𝑑𝑑𝑑𝑑 is called the total differential of the function
𝑧𝑧 = 𝑧𝑧(𝑥𝑥, 𝑦𝑦)
• The quantities like 𝜕𝜕𝜕𝜕/𝜕𝜕𝜕𝜕 𝑦𝑦 , i.e. the partial
derivative of 𝑧𝑧 with respect to 𝑥𝑥 holding 𝑦𝑦
constant, are the slopes of the meadow in
particular directions.
Mathematical Theorems
• Exact and Inexact Differentials: Suppose we are
given a differential expression of the form:
𝛿𝛿𝛿𝛿 𝑥𝑥, 𝑦𝑦 = 𝑀𝑀 𝑥𝑥, 𝑦𝑦 𝑑𝑑𝑑𝑑 + 𝑁𝑁 𝑥𝑥, 𝑦𝑦 𝑑𝑑𝑑𝑑
• Is this expression an exact differential?
• That is, can it be obtained from a function
𝑓𝑓(𝑥𝑥, 𝑦𝑦)?
• Answer: The answer to the above question is not
always yes.
• Suppose that, such a function does actually exist.
Mathematical Theorems
• If such a function 𝑓𝑓(𝑥𝑥, 𝑦𝑦) exists, then we have
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑦𝑦 𝜕𝜕𝜕𝜕 𝑥𝑥
• Thus comparing the above two, and since the changes in
the independent variables 𝑥𝑥 and 𝑦𝑦 are arbitrary (i.e.
independent of each other) we must have:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
= 𝑀𝑀(𝑥𝑥, 𝑦𝑦) and = 𝑁𝑁(𝑥𝑥, 𝑦𝑦)
𝜕𝜕𝜕𝜕 𝑦𝑦 𝜕𝜕𝜕𝜕 𝑥𝑥
• Hence, a differential of the form
𝛿𝛿𝛿𝛿 𝑥𝑥, 𝑦𝑦 = 𝑀𝑀 𝑥𝑥, 𝑦𝑦 𝑑𝑑𝑑𝑑 + 𝑁𝑁 𝑥𝑥, 𝑦𝑦 𝑑𝑑𝑑𝑑
will be an exact differential if the above conditions are true.
Mathematical Theorems
• However, if any two (arbitrary) functions 𝑀𝑀(𝑥𝑥, 𝑦𝑦)
and 𝑁𝑁 𝑥𝑥, 𝑦𝑦 are combined in such a way that:
𝛿𝛿𝐺𝐺 𝑥𝑥, 𝑦𝑦 = 𝑀𝑀 𝑥𝑥, 𝑦𝑦 𝑑𝑑𝑑𝑑 + 𝑁𝑁 𝑥𝑥, 𝑦𝑦 𝑑𝑑𝑑𝑑
in general, it is unlikely that the combination will
be an exact differential.
• Condition for Exact Differential:
• A differential 𝑑𝑑𝑑𝑑 = 𝑀𝑀 𝑥𝑥, 𝑦𝑦 𝑑𝑑𝑑𝑑 + 𝑁𝑁 𝑥𝑥, 𝑦𝑦 𝑑𝑑𝑑𝑑 is
exact, if any of the following statements are
satisfied:
Mathematical Theorems
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
• (a) 𝑀𝑀 𝑥𝑥, 𝑦𝑦 = and 𝑁𝑁 𝑥𝑥, 𝑦𝑦 =
𝜕𝜕𝜕𝜕 𝑦𝑦 𝜕𝜕𝜕𝜕 𝑥𝑥
𝐵𝐵
• (b) Its integral ∫𝐴𝐴 𝑑𝑑𝑑𝑑 is path independent, i.e.
𝐵𝐵
� 𝑑𝑑𝑑𝑑 = 𝑓𝑓𝐵𝐵 − 𝑓𝑓𝐴𝐴
𝐴𝐴
• (c) Hence, the integral along a closed contour is zero, i.e.
� 𝑑𝑑𝑑𝑑 = 0

𝜕𝜕𝜕𝜕 𝑥𝑥,𝑦𝑦 𝜕𝜕𝜕𝜕 𝑥𝑥,𝑦𝑦


• (d) =
𝜕𝜕𝜕𝜕 𝑥𝑥 𝜕𝜕𝜕𝜕 𝑦𝑦
Mathematical Theorems
• Proof: (a) If we move in the 𝑥𝑥𝑥𝑥 − plane from an initial point
𝑖𝑖 ≡ (𝑥𝑥𝑖𝑖 , 𝑦𝑦𝑖𝑖 ) to a final point 𝑗𝑗 ≡ (𝑥𝑥𝑗𝑗 , 𝑦𝑦𝑗𝑗 ), then the integral from
𝑗𝑗 𝑗𝑗
𝑖𝑖 to 𝑗𝑗 is: ∫𝑖𝑖 𝑑𝑑𝑑𝑑 = ∫𝑖𝑖 (𝑀𝑀𝑀𝑀𝑀𝑀 + 𝑁𝑁 𝑑𝑑𝑑𝑑)
• However, from the analogy of the displacement along the
meadow, and since 𝑑𝑑𝑑𝑑 and 𝑑𝑑𝑑𝑑 are independent of each other
,we get, if 𝑑𝑑𝑑𝑑 is exact, then we can write it as:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑦𝑦 𝜕𝜕𝜕𝜕 𝑥𝑥
• Hence, comparing the two expressions for 𝑑𝑑𝑑𝑑, we have:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑀𝑀 𝑥𝑥, 𝑦𝑦 = and 𝑁𝑁 𝑥𝑥, 𝑦𝑦 =
𝜕𝜕𝜕𝜕 𝑦𝑦 𝜕𝜕𝜕𝜕 𝑥𝑥
Proved (a).
Mathematical Theorems
• Proof: (b) If 𝑑𝑑𝑑𝑑 = 𝑀𝑀 𝑥𝑥, 𝑦𝑦 𝑑𝑑𝑑𝑑 + 𝑁𝑁 𝑥𝑥, 𝑦𝑦 𝑑𝑑𝑑𝑑 is an
exact differential, we get, upon integration
𝑗𝑗 𝑗𝑗
Δ𝑓𝑓 = 𝑓𝑓𝑗𝑗 − 𝑓𝑓𝑖𝑖 = � 𝑑𝑑𝑑𝑑 = � (𝑀𝑀𝑀𝑀𝑀𝑀 + 𝑁𝑁 𝑑𝑑𝑑𝑑)
𝑖𝑖 𝑖𝑖
• Since the difference on the left-hand side
depends only on the initial and final points, the
integral on the right-hand side can only depend
on these points as well.
𝑗𝑗
• Hence, the integral ∫𝑖𝑖 𝑑𝑑𝑑𝑑 is path independent.
Proved (b).
Mathematical Theorems
𝑗𝑗
• Proof: (c) Since the integral is path∫𝑖𝑖 𝑑𝑑𝑑𝑑
independent and depends only on the initial and
final points, we have trivially, ∮ 𝑑𝑑𝑑𝑑 = 0
• Proof: (d) We have, if 𝑑𝑑𝑑𝑑 = 𝑀𝑀 𝑥𝑥, 𝑦𝑦 𝑑𝑑𝑑𝑑 +
𝑁𝑁 𝑥𝑥, 𝑦𝑦 𝑑𝑑𝑑𝑑 is an exact differential:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑀𝑀 𝑥𝑥, 𝑦𝑦 = and𝑁𝑁 𝑥𝑥, 𝑦𝑦 =
𝜕𝜕𝜕𝜕 𝑦𝑦 𝜕𝜕𝜕𝜕 𝑥𝑥
𝜕𝜕2 𝑓𝑓(𝑥𝑥,𝑦𝑦) 𝜕𝜕𝜕𝜕 𝜕𝜕2 𝑓𝑓(𝑥𝑥,𝑦𝑦) 𝜕𝜕𝑁𝑁
• Then, = and =
𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑥𝑥 𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑦𝑦
Mathematical Theorems
• Since, in general for smooth enough functions,
𝜕𝜕2 𝑓𝑓(𝑥𝑥,𝑦𝑦) 𝜕𝜕2 𝑓𝑓(𝑥𝑥,𝑦𝑦)
=
𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕
we have,
𝜕𝜕𝜕𝜕 𝜕𝜕𝑁𝑁
=
𝜕𝜕𝜕𝜕 𝑥𝑥
𝜕𝜕𝜕𝜕 𝑦𝑦 Proved.
• Note: The exact conditions for the equality of the
mixed derivatives 𝑓𝑓𝑥𝑥𝑥𝑥 = 𝑓𝑓𝑦𝑦𝑦𝑦 are
• all of 𝑓𝑓𝑥𝑥 , 𝑓𝑓𝑦𝑦 , 𝑓𝑓𝑥𝑥𝑥𝑥 and 𝑓𝑓𝑦𝑦𝑦𝑦 exits and are continuous.
Mathematical Theorems
• Heuristic Proof of Equality of Mixed Partial
Derivatives:
• Consider 𝑧𝑧 = 𝑧𝑧(𝑥𝑥, 𝑦𝑦) as in the example of a
meadow.
• A good or smooth nature of the function 𝑧𝑧(𝑥𝑥, 𝑦𝑦)
is that its mixed partial derivatives are equal.
• Let us calculate the difference 𝛿𝛿𝛿𝛿 in the heights of
A and C in the figure to follow.
• We can go from A to C via B or via D, and 𝛿𝛿𝛿𝛿 is
route-independent.
Mathematical Theorems
• Then to first order:
(𝐴𝐴) (𝐵𝐵) (𝐴𝐴) (𝐷𝐷)
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝛿𝛿𝛿𝛿 = 𝛿𝛿𝛿𝛿 + 𝛿𝛿𝑦𝑦 = 𝛿𝛿𝑦𝑦 + 𝛿𝛿𝛿𝛿
𝜕𝜕𝜕𝜕 𝑦𝑦 𝜕𝜕𝑦𝑦 𝑥𝑥
𝜕𝜕𝑦𝑦 𝑥𝑥
𝜕𝜕𝜕𝜕 𝑦𝑦
• Here the superscript (A) means “evaluated at A”.
• Divide both sides by 𝛿𝛿𝛿𝛿𝛿𝛿𝛿𝛿:
𝜕𝜕𝜕𝜕 (𝐵𝐵) 𝜕𝜕𝜕𝜕 (𝐴𝐴) 𝜕𝜕𝜕𝜕 (𝐷𝐷) 𝜕𝜕𝜕𝜕 (𝐴𝐴)
− −
𝜕𝜕𝜕𝜕 𝑥𝑥 𝜕𝜕𝜕𝜕 𝑥𝑥 𝜕𝜕𝜕𝜕 𝑦𝑦 𝜕𝜕𝜕𝜕 𝑦𝑦
=
𝛿𝛿𝛿𝛿 𝛿𝛿𝛿𝛿
• Hence, in the infinitesimal limit:
𝜕𝜕 2 𝑧𝑧 𝜕𝜕 2 𝑧𝑧
=
𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕 𝜕𝜕𝑥𝑥𝜕𝜕𝑦𝑦

Proved.
Problems
• Example-1: Consider the differential
𝑑𝑑𝑑𝑑 = 𝐾𝐾1 𝑦𝑦 𝑑𝑑𝑑𝑑 + 𝐾𝐾2 𝑥𝑥 𝑑𝑑𝑑𝑑
where 𝐾𝐾1 , 𝐾𝐾2 are constants. Examine whether
𝑑𝑑𝑑𝑑 is an exact differential by integration over a
closed rectangular path in the 𝑥𝑥𝑥𝑥 −plane.
• Solution-1: Consider
a rectangular path
as shown below:
Problems
• Solution-1 (contd.): We evaluate the integral of 𝑑𝑑𝑑𝑑
along two paths between two points 𝐴𝐴 and 𝐵𝐵 as
shown.
(1) 𝑥𝑥𝐵𝐵 𝑦𝑦𝐵𝐵
• 𝑊𝑊 = 𝐾𝐾1 ∫𝑥𝑥 𝑑𝑑𝑑𝑑 𝑦𝑦𝐴𝐴+0+0+ 𝐾𝐾2 ∫𝑦𝑦 𝑑𝑑𝑑𝑑 𝑥𝑥𝐵𝐵
𝐴𝐴 𝐴𝐴
= 𝐾𝐾1 𝑦𝑦𝐴𝐴 𝑥𝑥𝐵𝐵 − 𝑥𝑥𝐴𝐴 + 𝐾𝐾2 𝑥𝑥𝐵𝐵 (𝑦𝑦𝐵𝐵 − 𝑦𝑦𝐴𝐴 )
(1) 𝑦𝑦𝐵𝐵 𝑥𝑥𝐵𝐵
• 𝑊𝑊 =0+ 𝐾𝐾2 ∫𝑦𝑦 𝑑𝑑𝑑𝑑 𝑥𝑥𝐴𝐴 + 𝐾𝐾1 ∫𝑥𝑥 𝑑𝑑𝑑𝑑 𝑦𝑦𝐵𝐵 +0
𝐴𝐴 𝐴𝐴
= 𝐾𝐾1 𝑦𝑦𝐵𝐵 𝑥𝑥𝐵𝐵 − 𝑥𝑥𝐴𝐴 + 𝐾𝐾2 𝑥𝑥𝐴𝐴 (𝑦𝑦𝐵𝐵 − 𝑦𝑦𝐴𝐴 )
• Hence in general, 𝑊𝑊 (1) ≠ 𝑊𝑊 (2)
• Comparing with the force 𝐹𝐹 and displacement along
the path, we see that the integrals will be work done.
Problems
• Thus if we start at point 𝐴𝐴, the kinetic energy at point 𝐵𝐵 will
depend on the path taken, since the work done is path-
dependent.
• The work done will be path independent if
𝑊𝑊 (1) − 𝑊𝑊 2
= � 𝑑𝑑𝑑𝑑 = (𝐾𝐾2 − 𝐾𝐾1 )(𝑥𝑥𝐵𝐵 − 𝑥𝑥𝐴𝐴 )(𝑦𝑦𝐵𝐵 − 𝑦𝑦𝐴𝐴 )

becomes zero.
• In fact, if 𝐾𝐾1 = 𝐾𝐾2 , the work would be path-independent,
and would depend only on the endpoints.
• Explicitly then, 𝐹𝐹 = 𝐾𝐾1 𝑥𝑥𝑥𝑥 . However, if 𝐾𝐾1 ≠ 𝐾𝐾2 , the
differential is inexact.
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Dependent and Independent
Variables
• Single-Component System: A single-component system is
specified by three state variables:
𝑃𝑃, 𝑉𝑉, 𝑇𝑇 or 𝑆𝑆, 𝑉𝑉, 𝑁𝑁 or 𝑈𝑈, 𝐻𝐻, 𝑇𝑇
or any three of 𝑃𝑃, 𝑉𝑉, 𝑇𝑇, 𝑆𝑆, 𝑁𝑁, 𝑈𝑈 etc.
• For example, we may have: 𝑈𝑈 = 𝑈𝑈(𝑆𝑆, 𝑉𝑉, 𝑁𝑁) or 𝑈𝑈 =
𝑈𝑈(𝑃𝑃, 𝑉𝑉, 𝑇𝑇), etc.
• In many systems, the number of particles 𝑁𝑁 is held
constant.
• This is called a constraint (i.e. an imposed condition).
• This makes the number of independent variables to two,
reducing it from three for a single-component hydrostatic
system.
Dependent and Independent
Variables
• Chemical Potential: We define the chemical potential
as the energy needed to increase the number of
particles of a system by one unit if everything else is
held constant.
• Introducing the possibility of change of the number of
particles, the first law has to be modified to:
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑑𝑑𝑄𝑄𝑖𝑖𝑖𝑖 + 𝑑𝑑𝑊𝑊𝑜𝑜𝑜𝑜 + 𝜇𝜇 𝑑𝑑𝑑𝑑
⇒ 𝑑𝑑𝑑𝑑 = 𝑇𝑇 𝑑𝑑𝑑𝑑 − 𝑃𝑃 𝑑𝑑𝑑𝑑 + 𝜇𝜇 𝑑𝑑𝑑𝑑
• The chemical potential is:
𝜕𝜕𝜕𝜕
𝜇𝜇 =
𝜕𝜕𝜕𝜕 𝑆𝑆,𝑉𝑉
Dependent and Independent
Variables
• The interdependence of so many different
thermodynamic variables means that there are
several different but equivalent ways of writing
any expression.
• The redundancy among the various parameters
may at first seem like a great deal of trouble.
• But we can use this interdependence to our
advantage.
• To do this, we need to have the necessary
mathematical relationships at our hand.
A General Mathematical Theorem
• Consider a set of 𝑛𝑛 independent variables:
{𝑥𝑥1 , 𝑥𝑥2 , 𝑥𝑥3 , … , 𝑥𝑥𝑛𝑛 } (for example: in 𝑈𝑈 =
𝑈𝑈(𝑃𝑃, 𝑉𝑉, 𝑇𝑇), the three variables are 𝑃𝑃, 𝑉𝑉, 𝑇𝑇 , etc. )
• We may describe this state as a point in
𝑛𝑛 −dimensional space.
• Again, the same state of the system may be
expressed by a set of other 𝑛𝑛 number of
independent variables: {𝑦𝑦1 , 𝑦𝑦2 , 𝑦𝑦3 , … , 𝑦𝑦𝑛𝑛 } (for
example, in 𝑈𝑈 = 𝑈𝑈(𝑆𝑆, 𝑉𝑉, 𝑁𝑁) they are 𝑆𝑆, 𝑉𝑉, 𝑁𝑁 )
• Or by a set of other 𝑛𝑛 state variables:
{𝑧𝑧1 , 𝑧𝑧2 , 𝑧𝑧3 , … , 𝑧𝑧𝑛𝑛 }.
A General Mathematical Theorem
• These three sets of variables are sets of independent
variables, but there exists transformation among the
elements of the sets.
• Hence, each variable from one set may be expressed as
a function of the variables in the other set(s) e. g.
𝑥𝑥𝑖𝑖 = 𝐹𝐹( 𝑦𝑦𝑖𝑖 ) or 𝑥𝑥𝑖𝑖 = 𝐺𝐺( 𝑧𝑧𝑖𝑖 )
• Then by chain rule between the variables:
𝜕𝜕𝑥𝑥𝑖𝑖 𝜕𝜕𝑥𝑥𝑖𝑖 𝜕𝜕𝑦𝑦𝑗𝑗 𝜕𝜕𝑥𝑥𝑖𝑖 𝜕𝜕𝑦𝑦𝑗𝑗
=� =
𝜕𝜕𝑧𝑧𝑘𝑘 𝜕𝜕𝑦𝑦𝑗𝑗 𝜕𝜕𝑧𝑧𝑘𝑘 𝜕𝜕𝑦𝑦𝑗𝑗 𝜕𝜕𝑧𝑧𝑘𝑘
𝑗𝑗
where we have used the Einstein summation convention.
A General Mathematical Theorem
• This can be written as a matrix multiplication:
𝐴𝐴𝑖𝑖𝑖𝑖 = 𝐵𝐵𝑖𝑖𝑖𝑖 𝐶𝐶𝑗𝑗𝑗𝑗
where, 𝐴𝐴𝑖𝑖𝑖𝑖 = 𝜕𝜕𝑥𝑥𝑖𝑖 /𝜕𝜕𝑧𝑧𝑘𝑘 , 𝐵𝐵𝑖𝑖𝑗𝑗 = 𝜕𝜕𝑥𝑥𝑖𝑖 /𝜕𝜕𝑦𝑦𝑗𝑗 , 𝐶𝐶𝑗𝑗𝑘𝑘 = 𝜕𝜕𝑦𝑦𝑗𝑗 /𝜕𝜕𝑧𝑧𝑘𝑘
• Define the Jacobian determinant:
𝜕𝜕𝑥𝑥𝑖𝑖 𝜕𝜕 𝑥𝑥1 , 𝑥𝑥2 , … , 𝑥𝑥𝑛𝑛
det =
𝜕𝜕𝑧𝑧𝑘𝑘 𝜕𝜕 𝑧𝑧1 , 𝑧𝑧2 , … , 𝑧𝑧𝑛𝑛
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝜕𝜕 𝑥𝑥,𝑦𝑦 𝜕𝜕𝑥𝑥𝑖𝑖 𝜕𝜕𝜕𝜕 𝑣𝑣 𝜕𝜕𝜕𝜕 𝑢𝑢
• For example: = det = 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝜕𝜕 𝑢𝑢,𝑣𝑣 𝜕𝜕𝑢𝑢𝑘𝑘
𝜕𝜕𝜕𝜕 𝑣𝑣 𝜕𝜕𝜕𝜕 𝑢𝑢
A General Mathematical Theorem
• Similarly, we can write:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝜕𝜕𝜕𝜕 𝑣𝑣,𝑤𝑤
𝜕𝜕𝜕𝜕 𝑢𝑢,𝑤𝑤
𝜕𝜕𝜕𝜕 𝑢𝑢,𝑣𝑣
𝜕𝜕 𝑥𝑥, 𝑦𝑦, 𝑧𝑧 𝜕𝜕𝑥𝑥𝑖𝑖 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
= det =
𝜕𝜕 𝑢𝑢, 𝑣𝑣, 𝑤𝑤 𝜕𝜕𝑢𝑢𝑘𝑘 𝜕𝜕𝜕𝜕 𝑣𝑣,𝑤𝑤
𝜕𝜕𝜕𝜕 𝑢𝑢,𝑤𝑤
𝜕𝜕𝜕𝜕 𝑢𝑢,𝑣𝑣
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝜕𝜕𝜕𝜕 𝑣𝑣,𝑤𝑤
𝜕𝜕𝜕𝜕 𝑢𝑢,𝑤𝑤
𝜕𝜕𝜕𝜕 𝑢𝑢,𝑣𝑣
• Omitting the subscript, for clarity we get:
𝜕𝜕𝜕𝜕/𝜕𝜕𝜕𝜕 𝜕𝜕𝑥𝑥/𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕/𝜕𝜕𝑤𝑤
𝜕𝜕 𝑥𝑥, 𝑦𝑦, 𝑧𝑧
= 𝜕𝜕𝑦𝑦/𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕/𝜕𝜕𝑣𝑣 𝜕𝜕𝜕𝜕/𝜕𝜕𝑤𝑤
𝜕𝜕 𝑢𝑢, 𝑣𝑣, 𝑤𝑤
𝜕𝜕𝑧𝑧/𝜕𝜕𝜕𝜕 𝜕𝜕𝑧𝑧/𝜕𝜕𝜕𝜕 𝜕𝜕𝑧𝑧/𝜕𝜕𝜕𝜕
A General Mathematical Theorem
• From the properties of the determinant, we get, if 𝐴𝐴 = 𝐵𝐵𝐵𝐵
then, det 𝐴𝐴 = det(𝐵𝐵) det(𝐶𝐶)
• Hence, we get:
𝜕𝜕 𝑥𝑥1 ,𝑥𝑥2 ,…,𝑥𝑥𝑛𝑛 𝜕𝜕 𝑥𝑥1 ,𝑥𝑥2 ,…,𝑥𝑥𝑛𝑛 𝜕𝜕 𝑦𝑦1 ,𝑦𝑦2 ,…,𝑦𝑦𝑛𝑛
= .
𝜕𝜕 𝑧𝑧1 ,𝑧𝑧2 ,…,𝑧𝑧𝑛𝑛 𝜕𝜕 𝑦𝑦1 ,𝑦𝑦2 ,…,𝑦𝑦𝑛𝑛 𝜕𝜕 𝑧𝑧1 ,𝑧𝑧2 ,…,𝑧𝑧𝑛𝑛
• Explicitly for the case of 𝑛𝑛 = 2 we get,
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝜕𝜕 𝑥𝑥, 𝑦𝑦 𝜕𝜕𝜕𝜕 𝑣𝑣
𝜕𝜕𝜕𝜕 𝑢𝑢 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
= = −
𝜕𝜕 𝑢𝑢, 𝑣𝑣 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑣𝑣
𝜕𝜕𝜕𝜕 𝑢𝑢
𝜕𝜕𝜕𝜕 𝑣𝑣
𝜕𝜕𝜕𝜕 𝑢𝑢
𝜕𝜕𝜕𝜕 𝑣𝑣
𝜕𝜕𝜕𝜕 𝑢𝑢
𝜕𝜕 𝑢𝑢,𝑣𝑣 𝜕𝜕𝑢𝑢 𝜕𝜕𝑣𝑣 𝜕𝜕𝑣𝑣 𝜕𝜕𝑢𝑢
• Similarly, = −
𝜕𝜕 𝑟𝑟,𝑠𝑠 𝜕𝜕𝑟𝑟 𝑠𝑠 𝜕𝜕𝑠𝑠 𝑟𝑟 𝜕𝜕𝑟𝑟 𝑠𝑠 𝜕𝜕𝑠𝑠 𝑟𝑟
A General Mathematical Theorem
𝜕𝜕 𝑥𝑥,𝑦𝑦 𝜕𝜕𝑥𝑥 𝜕𝜕𝑦𝑦 𝜕𝜕𝑦𝑦 𝜕𝜕𝑥𝑥
• Also, = −
𝜕𝜕 𝑟𝑟,𝑠𝑠 𝜕𝜕𝜕𝜕 𝑠𝑠 𝜕𝜕𝜕𝜕 𝑟𝑟 𝜕𝜕𝜕𝜕 𝑠𝑠 𝜕𝜕𝜕𝜕 𝑟𝑟
𝜕𝜕 𝑥𝑥,𝑦𝑦 𝜕𝜕 𝑢𝑢,𝑣𝑣 𝜕𝜕 𝑥𝑥,𝑦𝑦
• Hence as an example: . =
𝜕𝜕 𝑢𝑢,𝑣𝑣 𝜕𝜕 𝑟𝑟,𝑠𝑠 𝜕𝜕 𝑟𝑟,𝑠𝑠
𝜕𝜕𝑥𝑥𝑖𝑖
• Also recall that: = 𝛿𝛿𝑖𝑖𝑖𝑖
𝜕𝜕𝑥𝑥𝑘𝑘
• From this simple mathematics follows several
very useful results:
A General Mathematical Theorem
𝜕𝜕𝜕𝜕 𝜕𝜕𝑢𝑢
• Relation-1: = 1/
𝜕𝜕𝜕𝜕 𝑦𝑦 𝜕𝜕𝑥𝑥 𝑦𝑦

• Proof: From det 𝐴𝐴 . det 𝐴𝐴−1 = det(𝐴𝐴. 𝐴𝐴−1 ) = det 𝐼𝐼 =


𝜕𝜕(𝑥𝑥,𝑦𝑦) 𝜕𝜕 𝑢𝑢,𝑣𝑣 −1
1 we get, =
𝜕𝜕(𝑢𝑢,𝑣𝑣) 𝜕𝜕 𝑥𝑥,𝑦𝑦
• Now let, 𝑣𝑣 = 𝑦𝑦, which gives,
𝜕𝜕(𝑥𝑥, 𝑦𝑦) 𝜕𝜕(𝑥𝑥, 𝑦𝑦) 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
= = −
𝜕𝜕(𝑢𝑢, 𝑣𝑣) 𝜕𝜕(𝑢𝑢, 𝑦𝑦) 𝜕𝜕𝜕𝜕 𝑦𝑦
𝜕𝜕𝜕𝜕 𝑢𝑢
𝜕𝜕𝜕𝜕 𝑦𝑦
𝜕𝜕𝜕𝜕 𝑢𝑢
𝜕𝜕(𝑥𝑥,𝑦𝑦) 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
• ⇒ = . 1 − 0. =
𝜕𝜕(𝑢𝑢,𝑦𝑦) 𝜕𝜕𝜕𝜕 𝑦𝑦 𝜕𝜕𝜕𝜕 𝑢𝑢 𝜕𝜕𝜕𝜕 𝑦𝑦
A General Mathematical Theorem
𝜕𝜕 𝑢𝑢,𝑦𝑦 𝜕𝜕𝜕𝜕
• Similarly, =
𝜕𝜕 𝑥𝑥,𝑦𝑦 𝜕𝜕𝜕𝜕 𝑦𝑦

𝜕𝜕(𝑥𝑥,𝑦𝑦) 𝜕𝜕𝜕𝜕 𝜕𝜕 𝑢𝑢,𝑦𝑦 −1 𝜕𝜕𝜕𝜕


• ⇒ = = = 1/
𝜕𝜕(𝑢𝑢,𝑦𝑦) 𝜕𝜕𝜕𝜕 𝑦𝑦 𝜕𝜕 𝑥𝑥,𝑦𝑦 𝜕𝜕𝜕𝜕 𝑦𝑦
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
• Hence, = 1/ Proved.
𝜕𝜕𝜕𝜕 𝑦𝑦 𝜕𝜕𝜕𝜕 𝑦𝑦
A General Mathematical Theorem
• Relation-2:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
= −1
𝜕𝜕𝜕𝜕 𝑢𝑢
𝜕𝜕𝜕𝜕 𝑥𝑥
𝜕𝜕𝜕𝜕 𝑦𝑦
• Proof: We also get,
𝜕𝜕(𝑥𝑥, 𝑦𝑦) 𝜕𝜕𝜕𝜕 𝜕𝜕(𝑥𝑥, 𝑦𝑦) 𝜕𝜕(𝑥𝑥, 𝑢𝑢)
= = .
𝜕𝜕(𝑢𝑢, 𝑦𝑦) 𝜕𝜕𝜕𝜕 𝑦𝑦
𝜕𝜕(𝑥𝑥, 𝑢𝑢) 𝜕𝜕(𝑢𝑢, 𝑦𝑦)
𝜕𝜕(𝑥𝑥,𝑦𝑦) 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
• But, = −
𝜕𝜕(𝑥𝑥,𝑢𝑢) 𝜕𝜕𝑥𝑥 𝑢𝑢 𝜕𝜕𝑢𝑢 𝑥𝑥 𝜕𝜕𝑥𝑥 𝑢𝑢 𝜕𝜕𝑢𝑢 𝑥𝑥
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
= 1. − .0 =
𝜕𝜕𝜕𝜕 𝑥𝑥
𝜕𝜕𝜕𝜕 𝑢𝑢
𝜕𝜕𝜕𝜕 𝑥𝑥
A General Mathematical Theorem
𝜕𝜕(𝑥𝑥,𝑢𝑢) 𝜕𝜕𝜕𝜕 𝜕𝜕𝑢𝑢 𝜕𝜕𝑢𝑢 𝜕𝜕𝜕𝜕
• But, = −
𝜕𝜕(𝑢𝑢,𝑦𝑦) 𝜕𝜕𝑢𝑢 𝑦𝑦 𝜕𝜕𝑦𝑦 𝑢𝑢 𝜕𝜕𝑢𝑢 𝑦𝑦 𝜕𝜕𝑦𝑦 𝑢𝑢
𝜕𝜕𝑥𝑥 𝜕𝜕𝑥𝑥 𝜕𝜕𝜕𝜕
= . 0 − 1. =−
𝜕𝜕𝜕𝜕 𝑦𝑦
𝜕𝜕𝑦𝑦 𝑢𝑢
𝜕𝜕𝜕𝜕 𝑢𝑢
• Hence,
𝜕𝜕(𝑥𝑥,𝑦𝑦) 𝜕𝜕𝜕𝜕 𝜕𝜕(𝑥𝑥,𝑦𝑦) 𝜕𝜕(𝑥𝑥,𝑢𝑢) 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
= = . =−
𝜕𝜕(𝑢𝑢,𝑦𝑦) 𝜕𝜕𝜕𝜕 𝑦𝑦 𝜕𝜕(𝑥𝑥,𝑢𝑢) 𝜕𝜕(𝑢𝑢,𝑦𝑦) 𝜕𝜕𝜕𝜕 𝑥𝑥 𝜕𝜕𝜕𝜕 𝑢𝑢

𝜕𝜕𝜕𝜕 𝜕𝜕𝑦𝑦 𝜕𝜕𝜕𝜕


• Hence, = −1
𝜕𝜕𝜕𝜕 𝑢𝑢 𝜕𝜕𝑢𝑢 𝑥𝑥 𝜕𝜕𝜕𝜕 𝑦𝑦
A General Mathematical Theorem
• Relation-3:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
=
𝜕𝜕𝜕𝜕 𝑣𝑣 𝜕𝜕𝜕𝜕 𝑣𝑣 𝜕𝜕𝜕𝜕 𝑣𝑣
• Proof: We have,
𝜕𝜕(𝑥𝑥, 𝑣𝑣) 𝜕𝜕(𝑥𝑥, 𝑣𝑣) 𝜕𝜕(𝑦𝑦, 𝑣𝑣)
= .
𝜕𝜕(𝑢𝑢, 𝑣𝑣) 𝜕𝜕(𝑦𝑦, 𝑣𝑣) 𝜕𝜕(𝑢𝑢, 𝑣𝑣)
• Which says:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝑦𝑦
=
𝜕𝜕𝜕𝜕 𝑣𝑣 𝜕𝜕𝑦𝑦 𝑣𝑣 𝜕𝜕𝜕𝜕 𝑣𝑣
which is simply the chain rule of partial differentiation.
A General Mathematical Theorem
• Relation-4:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
= −
𝜕𝜕𝜕𝜕 𝑦𝑦
𝜕𝜕𝜕𝜕 𝑣𝑣
𝜕𝜕𝜕𝜕 𝑢𝑢
𝜕𝜕𝜕𝜕 𝑣𝑣
• Proof: We have,
𝜕𝜕(𝑥𝑥, 𝑦𝑦) 𝜕𝜕(𝑥𝑥, 𝑦𝑦) 𝜕𝜕(𝑢𝑢, 𝑣𝑣)
= .
𝜕𝜕(𝑢𝑢, 𝑦𝑦) 𝜕𝜕(𝑢𝑢, 𝑣𝑣) 𝜕𝜕(𝑢𝑢, 𝑦𝑦)
𝜕𝜕(𝑥𝑥,𝑦𝑦) 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
• But, 𝜕𝜕(𝑢𝑢,𝑦𝑦)
=
𝜕𝜕𝜕𝜕 𝑦𝑦 𝜕𝜕𝑦𝑦 𝑢𝑢

𝜕𝜕𝜕𝜕 𝑦𝑦 𝜕𝜕𝑦𝑦 𝑢𝑢
=
𝜕𝜕𝜕𝜕 𝑦𝑦
𝜕𝜕(𝑥𝑥,𝑦𝑦) 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
• Also, = −
𝜕𝜕(𝑢𝑢,𝑣𝑣) 𝜕𝜕𝜕𝜕 𝑣𝑣 𝜕𝜕𝑣𝑣 𝑢𝑢 𝜕𝜕𝜕𝜕 𝑣𝑣 𝜕𝜕𝑣𝑣 𝑢𝑢
A General Mathematical Theorem
𝜕𝜕(𝑢𝑢,𝑣𝑣) 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
• And, = − =
𝜕𝜕(𝑢𝑢,𝑦𝑦) 𝜕𝜕𝜕𝜕 𝑦𝑦 𝜕𝜕𝜕𝜕 𝑢𝑢 𝜕𝜕𝜕𝜕 𝑦𝑦 𝜕𝜕𝜕𝜕 𝑢𝑢 𝜕𝜕𝜕𝜕 𝑢𝑢
• Hence,
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
= −
𝜕𝜕𝜕𝜕 𝑦𝑦 𝜕𝜕𝜕𝜕 𝑢𝑢
𝜕𝜕𝜕𝜕 𝑣𝑣
𝜕𝜕𝜕𝜕 𝑢𝑢
𝜕𝜕𝜕𝜕 𝑣𝑣
𝜕𝜕𝜕𝜕 𝑢𝑢
• Hence we get,
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
= −
𝜕𝜕𝜕𝜕 𝑦𝑦
𝜕𝜕𝜕𝜕 𝑣𝑣
𝜕𝜕𝑦𝑦 𝑢𝑢
𝜕𝜕𝜕𝜕 𝑣𝑣
A General Mathematical Theorem
• Example-1: Consider the transformation between
plane polar and 2D Cartesian coordinates.
• 𝑥𝑥 = 𝑟𝑟 cos 𝜃𝜃, 𝑦𝑦 = 𝑟𝑟 sin 𝜃𝜃, 𝑟𝑟 = 𝑥𝑥 2 + 𝑦𝑦 2 ,
𝜃𝜃 = tan−1 (𝑦𝑦/𝑥𝑥)
𝜕𝜕𝜕𝜕
• Then, = cos 𝜃𝜃
𝜕𝜕𝜕𝜕
• Naively, we would expect that
the inverse of this would be:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 1
= 1/ = = sec 𝜃𝜃 ?
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 cos 𝜃𝜃
• Is it true?
A General Mathematical Theorem
• Example-1 (Contd.):
• The answer is NO.
𝜕𝜕𝜕𝜕 𝜕𝜕 1 2𝑥𝑥 𝑥𝑥
• = 𝑥𝑥 2 + 𝑦𝑦 2 = = = cos 𝜃𝜃 !
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 2 𝑥𝑥 2 +𝑦𝑦 2 𝑟𝑟
• Where is the catch?
• The answer lies in the choice of the variable held
constant in partial derivatives.
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
• Note that: we have = 1/ where in ,
𝜕𝜕𝜕𝜕 𝑦𝑦 𝜕𝜕𝜕𝜕 𝑦𝑦
both the derivatives, the same variable is held
constant.
A General Mathematical Theorem
• Take, 𝑢𝑢 = 𝑟𝑟 and 𝑣𝑣 = 𝜃𝜃.
𝜕𝜕𝑥𝑥 𝜕𝜕
• Then, = 𝑟𝑟 cos 𝜃𝜃� = cos 𝜃𝜃
𝜕𝜕𝑟𝑟 𝜃𝜃 𝜕𝜕𝜕𝜕 𝜃𝜃
𝜕𝜕
𝜕𝜕𝜕𝜕 𝜕𝜕 1 2𝑥𝑥+2𝑦𝑦 𝑟𝑟 sin 𝜃𝜃
• But, = 𝑥𝑥 2 + 𝑦𝑦 2 � = 𝜕𝜕𝜕𝜕
𝜕𝜕𝜕𝜕 𝜃𝜃 𝜕𝜕𝜕𝜕 𝜃𝜃 2 𝑥𝑥 2 +𝑦𝑦 2
𝜕𝜕𝜕𝜕 𝑥𝑥 𝑦𝑦 sin 𝜃𝜃 𝜕𝜕𝜕𝜕
⇒ = +
𝜕𝜕𝜕𝜕 𝜃𝜃
𝑟𝑟 𝑟𝑟 𝜕𝜕𝜕𝜕 𝜃𝜃
𝜕𝜕𝜕𝜕 2
𝜕𝜕𝜕𝜕
⇒ 1 − sin 𝜃𝜃 = cos 𝜃𝜃 ⇒ = sec 𝜃𝜃
𝜕𝜕𝜕𝜕 𝜃𝜃 𝜕𝜕𝜕𝜕 𝜃𝜃
as expected !!
A General Mathematical Theorem
𝜕𝜕𝜕𝜕 𝜕𝜕 1 2𝑥𝑥 𝑥𝑥
• Again, = 𝑥𝑥 2 + 𝑦𝑦 2 = = = cos 𝜃𝜃
𝜕𝜕𝜕𝜕 𝑦𝑦 𝜕𝜕𝜕𝜕 2 𝑥𝑥 2 +𝑦𝑦 2 𝑟𝑟
𝜕𝜕𝑥𝑥 𝜕𝜕 𝜕𝜕𝜕𝜕
• = 𝑟𝑟 cos 𝜃𝜃� = cos 𝜃𝜃 − 𝑟𝑟 sin 𝜃𝜃
𝜕𝜕𝑟𝑟 𝑦𝑦 𝜕𝜕𝜕𝜕 𝑦𝑦 𝜕𝜕𝜕𝜕 𝑦𝑦
𝜕𝜕𝜕𝜕 𝜕𝜕 −1 𝑦𝑦 1
(−𝑦𝑦) 𝜕𝜕𝜕𝜕
• But, = tan � = 𝑦𝑦2 𝑥𝑥 2
𝜕𝜕𝜕𝜕 𝑦𝑦 𝜕𝜕𝜕𝜕 𝑥𝑥 𝑦𝑦 1+ 2 𝜕𝜕𝜕𝜕 𝑦𝑦
𝑥𝑥
𝜕𝜕𝜕𝜕 𝑦𝑦 𝜕𝜕𝜕𝜕 sin 𝜃𝜃 𝜕𝜕𝜕𝜕
⇒ =− 2 =−
𝜕𝜕𝜕𝜕 𝑦𝑦
𝑟𝑟 𝜕𝜕𝜕𝜕 𝑦𝑦
𝑟𝑟 𝜕𝜕𝜕𝜕 𝑦𝑦
𝜕𝜕𝜕𝜕 2 𝜕𝜕𝜕𝜕
• Hence, 1 − sin 𝜃𝜃 = cos 𝜃𝜃 ⇒ = sec 𝜃𝜃 !!
𝜕𝜕𝜕𝜕 𝑦𝑦 𝜕𝜕𝜕𝜕 𝑦𝑦
Changing Variables
• Suppose that we want to determine changes in some
property 𝑋𝑋 for a closed system.
• The constraint 𝑑𝑑𝑑𝑑 = 0 reduces the number of
independent variables from three to two.
• Suppose we have at hand:
A. Thermometer
B. Pressure Gauge.
• Then we would like to have an expression like:
𝑑𝑑 𝑋𝑋 = 𝐴𝐴 𝑑𝑑𝑑𝑑 + 𝐵𝐵 𝑑𝑑𝑑𝑑
or, Δ𝑥𝑥 = 𝐴𝐴 Δ𝑇𝑇 + 𝐵𝐵 Δ𝑃𝑃 for finite changes.
• How do we get such an expression?
Changing Variables
• Imagine we have several variables 𝑣𝑣, 𝑤𝑤, 𝑥𝑥, 𝑦𝑦, . . . , of which
only two are independent.
• Suppose that we wish to know how one varies with two
others (say how 𝑥𝑥 varies with 𝑦𝑦 and 𝑧𝑧).
• Direct Approach: The most direct way of attacking this
problem is to write:
𝜕𝜕𝜕𝜕 𝜕𝜕𝑦𝑦
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑧𝑧 = 𝐴𝐴 𝑑𝑑𝑑𝑑 + 𝐵𝐵 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑧𝑧 𝜕𝜕𝑧𝑧 𝑦𝑦
• The partial derivatives may not be very useful as they
stand.
Changing Variables
• But if we are lucky or clever, we can convert them
into easily-measured properties, such as
pressure, temperature, heat capacities, etc.
• Indirect Approach: Suppose the partials of the
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
direct approach, e.g., and , are
𝜕𝜕𝜕𝜕 𝑧𝑧 𝜕𝜕𝜕𝜕 𝑦𝑦
inconvenient or not helpful for us.
• If we know how x varies with respect to any other
variables then we can use the partials of these
other variables instead.
Changing Variables
• Suppose that we do know how 𝑥𝑥 varies with 𝑣𝑣 and 𝑤𝑤,
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
i.e. we know and :
𝜕𝜕𝜕𝜕 𝑤𝑤 𝜕𝜕𝑤𝑤 𝑣𝑣
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑓𝑓 𝑑𝑑𝑑𝑑 + 𝑔𝑔 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑤𝑤
𝜕𝜕𝜕𝜕 𝑣𝑣
• How do we use these to find the dependence of 𝑥𝑥 on 𝑦𝑦
and 𝑧𝑧?
• We first write the derivatives 𝑑𝑑𝑑𝑑, 𝑑𝑑𝑑𝑑 each in terms of
𝑑𝑑𝑑𝑑, 𝑑𝑑𝑑𝑑 :
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝑤𝑤 𝜕𝜕𝑤𝑤
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑧𝑧 , 𝑑𝑑𝑤𝑤 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑧𝑧 𝜕𝜕𝑧𝑧 𝑦𝑦 𝜕𝜕𝜕𝜕 𝑧𝑧 𝜕𝜕𝜕𝜕 𝑦𝑦
Changing Variables
• Then we have: 𝑑𝑑𝑑𝑑 = 𝑓𝑓𝑓𝑓𝑓𝑓 + 𝑔𝑔 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑓𝑓 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑 + 𝑔𝑔 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑧𝑧 𝜕𝜕𝜕𝜕 𝑦𝑦
𝜕𝜕𝜕𝜕 𝑧𝑧
𝜕𝜕𝜕𝜕 𝑦𝑦
• Comparing with 𝑑𝑑𝑑𝑑 = 𝐴𝐴 𝑑𝑑𝑑𝑑 + 𝐵𝐵 𝑑𝑑𝑑𝑑 we get:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝐴𝐴 = 𝑓𝑓 + 𝑔𝑔
𝜕𝜕𝜕𝜕 𝑧𝑧 𝜕𝜕𝜕𝜕 𝑧𝑧
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝐵𝐵 = 𝑓𝑓 + 𝑔𝑔
𝜕𝜕𝜕𝜕 𝑦𝑦 𝜕𝜕𝜕𝜕 𝑦𝑦
• Although we now have the desired variables, 𝑦𝑦, 𝑧𝑧, the coefficients
are a mess.
• But we can use the rules for the change of variables to get the
desired expressions.
Why switch Variables?
• It is clear that, through transformations such as
the above, we can write thermodynamic
expressions in terms of virtually whichever set of
variables we wish.
• But why would we prefer one set of variables
over another?
• This question is easy to answer.
• It is clearly to our advantage to express the
properties of a system in terms of parameters
that are either (a) constrained, or (b) easy to
measure.
Why switch Variables?
• Example-2: Suppose that we are interested in some
property 𝑋𝑋 of a system undergoing a non-diffusive (i.e. 𝑁𝑁 is
not changing) process at atmospheric pressure .
• We also have a thermometer at our disposal.
• Clearly, 𝑑𝑑𝑑𝑑 = 0 and 𝑑𝑑𝑑𝑑 = 0.
• Then it is advantageous to express 𝑋𝑋 as a function of 𝑇𝑇, 𝑃𝑃, 𝑁𝑁
i.e.:
𝑑𝑑𝑑𝑑 = 𝐴𝐴 𝑑𝑑𝑑𝑑 + 𝐵𝐵 𝑑𝑑𝑑𝑑 + 𝐶𝐶 𝑑𝑑𝑑𝑑 = 𝐴𝐴 𝑑𝑑𝑑𝑑
where 𝑑𝑑𝑑𝑑 is easy to measure.
• The coefficient 𝐴𝐴 = 𝜕𝜕𝜕𝜕/𝜕𝜕𝜕𝜕 𝑃𝑃,𝑁𝑁 has some partial
derivatives, but we can write them in terms of easily
measurable quantities.
Why switch Variables?
• Example-3: Consider the adiabatic expansion and
cooling of a rising air mass that remains intact (in
atmospheric physics).
• Then clearly, 𝑑𝑑𝑑𝑑 = 0 (intact mass) , 𝑑𝑑𝑑𝑑 = 0
(adiabatic).
• It is them convenient to take a quantity 𝑋𝑋 as a
function of 𝑆𝑆, 𝑁𝑁 and another variable, such as the
pressure 𝑃𝑃:
𝑑𝑑𝑑𝑑 = 𝐴𝐴 𝑑𝑑𝑑𝑑 + 𝐵𝐵 𝑑𝑑𝑑𝑑 + 𝐶𝐶 𝑑𝑑𝑑𝑑 = 𝐴𝐴 𝑑𝑑𝑑𝑑
• We can express the partial derivative 𝐴𝐴 in terms
of easily measurable properties of the air-mass.
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Legendre Transformation
• The Legendre transforms appear in two places in
a standard undergraduate physics curriculum:
• (1) in classical mechanics when one switches
from Lagrangian to Hamiltonian dynamics, and
• (2) in thermodynamics to motivate the
connection between the internal energy,
enthalpy, and Gibbs and Helmholtz free energies,
etc.
Legendre Transformation
• Motivation: Consider the function 𝑦𝑦 = 𝑓𝑓(𝑥𝑥) ( i.e. the
mapping 𝑥𝑥 → 𝑦𝑦 = 𝑓𝑓(𝑥𝑥) ).
• This contains a lot of information. For example:
• A. It tells the 𝑦𝑦-value for different values of 𝑥𝑥.
• B. It also tells us the slope of the tangent to the graph
𝑦𝑦 = 𝑓𝑓(𝑥𝑥) at any point 𝑥𝑥 (as long as 𝑓𝑓 is differentiable).
• Sometimes that derivative (i.e. the slope) turns out to
be of so much interest, that one is tempted to use it as
an independent variable itself:
𝑝𝑝 = 𝑓𝑓 ′ (𝑥𝑥)
Legendre Transformation
• We might want to find a way to express the
information content of the function 𝑦𝑦 = 𝑓𝑓(𝑥𝑥)
using the derivative 𝑝𝑝 as the independent
variable.
• Of course, we would not want to lose any
information along the way.
• So how could we do that?
• We may try to solve 𝑝𝑝 = 𝑓𝑓 ′ 𝑥𝑥 = ℎ(𝑥𝑥) and find
𝑥𝑥 = ℎ−1 (𝑝𝑝) , where ℎ−1 denotes the inverse
function of 𝑓𝑓 ′ ( i.e. ℎ−1 ≡ 𝑓𝑓 ′ −1 )
Legendre Transformation
• The crucial question is: Does this new function
ℎ−1 contain the same information content as the
original function 𝑓𝑓?
• The answer is NO.
• The best way to see this is by a simple example.
1 2
• Take: 𝑦𝑦 = 𝑓𝑓 𝑥𝑥 = 𝑥𝑥 − 𝑥𝑥0
2
• Then the derivative, 𝑝𝑝 = 𝑓𝑓 ′ 𝑥𝑥 = 𝑦𝑦 ′ 𝑥𝑥 = 𝑥𝑥 − 𝑥𝑥0
• This can uniquely solved for 𝑥𝑥 as a function of 𝑝𝑝:
𝑥𝑥 = 𝑥𝑥0 + 𝑝𝑝 i.e. 𝑥𝑥 = 𝑥𝑥 𝑝𝑝 = ℎ−1 𝑝𝑝 = 𝑥𝑥0 + 𝑝𝑝.
Legendre Transformation
• If we insert it back to our original function 𝑦𝑦 = 𝑓𝑓 𝑥𝑥 =
𝑦𝑦(𝑥𝑥), we get:
1 −1
−1
𝑦𝑦 𝑥𝑥 = 𝑦𝑦(𝑥𝑥 𝑝𝑝 ) = 𝑦𝑦 ℎ 𝑝𝑝 = ℎ 𝑝𝑝 − 𝑥𝑥0 2
2
1 2
1 2
⇒ 𝑦𝑦 𝑥𝑥 = 𝑝𝑝 + 𝑥𝑥0 − 𝑥𝑥0 = 𝑝𝑝
2 2
• Thus the value of 𝑥𝑥0 has dropped out!
• Functions with different values of 𝑥𝑥0 would map to the
1 2
same final function 𝑝𝑝
2
• There is a way to solve this problem, but not for all
possible functions.
Legendre Transformation
• It turns out that a transformation from a function
𝑦𝑦 = 𝑓𝑓(𝑥𝑥) to a new function 𝑔𝑔(𝑝𝑝) where:
• (i) 𝑝𝑝 = 𝑓𝑓 ′ 𝑥𝑥 = 𝑦𝑦 ′ 𝑥𝑥 and
• (ii) no information is lost is possible, if and only if the
function 𝑦𝑦 = 𝑓𝑓(𝑥𝑥) is either convex or concave.
• Convex Function: A convex function of a single variable
𝑥𝑥 i.e. 𝑓𝑓(𝑥𝑥) is one for which 𝑓𝑓′′(𝑥𝑥) > 0 everywhere.
• Concave Function: A concave function of a single
variable 𝑥𝑥 i.e. 𝑓𝑓(𝑥𝑥) is one for which 𝑓𝑓 ′′ 𝑥𝑥 < 0
everywhere.
Legendre Transformation
• A convex function is also called a concave up
function.
Legendre Transformation
• The slope of a convex function is a
monotonically increasing function of the
independent variable.
Legendre Transformation
• Definition: The Legendre transform of a
convex function 𝑓𝑓(𝑥𝑥) is a function 𝑔𝑔 𝑝𝑝
defined by:
𝑔𝑔 𝑝𝑝 = 𝑚𝑚𝑚𝑚𝑚𝑚𝑖𝑖𝑥𝑥 [𝑝𝑝𝑝𝑝 − 𝑓𝑓(𝑥𝑥)]
where 𝑚𝑚𝑚𝑚𝑚𝑚𝑖𝑖𝑥𝑥 [𝑝𝑝𝑝𝑝 − 𝑓𝑓(𝑥𝑥)] implies that the
maximization of 𝑝𝑝𝑝𝑝 − 𝑓𝑓(𝑥𝑥) is done, with respect
to the variable 𝑥𝑥 while keeping 𝑝𝑝 fixed.
• For a concave function 𝑓𝑓(𝑥𝑥) , we have instead:
𝑔𝑔 𝑝𝑝 = 𝑚𝑚𝑖𝑖𝑖𝑖𝑖𝑖𝑥𝑥 [𝑝𝑝𝑝𝑝 − 𝑓𝑓(𝑥𝑥)]
Legendre Transformation
• Geometric Explanation: Consider a convex
function 𝑦𝑦 = 𝑓𝑓(𝑥𝑥).
• Let 𝑝𝑝 be a given real number, and consider the
line 𝑦𝑦 = 𝑝𝑝𝑝𝑝.

�(𝒑𝒑) − 𝒈𝒈 𝒑𝒑
𝒚𝒚 = 𝒑𝒑𝒙𝒙
�(𝒑𝒑))
= 𝒇𝒇(𝒙𝒙

�(𝒑𝒑)
𝒙𝒙
Legendre Transformation
• We maximize 𝑝𝑝𝑝𝑝 − 𝑓𝑓(𝑥𝑥) with respect to 𝑥𝑥 keeping 𝑝𝑝 fixed:
𝑑𝑑
𝑥𝑥𝑥𝑥 − 𝑓𝑓 𝑥𝑥 � =0
𝑑𝑑𝑑𝑑 𝑥𝑥=𝑥𝑥�, 𝑝𝑝−𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓𝑓
⇒ 𝑝𝑝 = 𝑓𝑓 ′ 𝑥𝑥 |𝑥𝑥� = 𝑓𝑓 ′ (𝑥𝑥�)
where, 𝑥𝑥� = 𝑥𝑥�(𝑝𝑝) is the point where 𝑥𝑥𝑥𝑥 − 𝑓𝑓 𝑥𝑥 becomes the
maximum.
𝑑𝑑 2 𝑑𝑑 2 𝑓𝑓 𝑥𝑥
• Note that, 𝑥𝑥𝑥𝑥 − 𝑓𝑓(𝑥𝑥) =− < 0 since 𝑓𝑓 was
𝑑𝑑𝑥𝑥 2 𝑑𝑑𝑥𝑥 2
assumed to be convex.
• Hence, at 𝑥𝑥 = 𝑥𝑥�, 𝑥𝑥𝑥𝑥 − 𝑓𝑓 𝑥𝑥 does become a maximum,
indeed.
Legendre Transformation
• Alternative Geometric Interpretation:
• Consider a convex function 𝑓𝑓(𝑥𝑥).
• Choose a value of 𝒙𝒙 which is represented by the
horizontal line-segment from the origin.
• Go up to the value of 𝒚𝒚 =
𝒇𝒇(𝒙𝒙)
the function curve 𝑓𝑓(𝑥𝑥).
• Draw the tangent to the
curve at that point. 𝒇𝒇(𝒙𝒙)
• The slope is 𝑝𝑝.
𝒙𝒙
Legendre Transformation
• Extend the tangent until it hits the 𝑦𝑦 −axis.
• The intercept is negative and equal to −𝑔𝑔(𝑝𝑝).
• Because the slope of the 𝒚𝒚 = Slope = 𝒑𝒑
𝒇𝒇(𝒙𝒙)
tangent is 𝑝𝑝 the length 𝒚𝒚 = 𝒑𝒑𝒑𝒑
−𝒈𝒈(𝒑𝒑)
of the dotted line is 𝑝𝑝𝑝𝑝.
• Hence, 𝒑𝒑𝒑𝒑 𝒇𝒇(𝒙𝒙)

𝑝𝑝𝑝𝑝 = 𝑓𝑓 𝑥𝑥 + 𝑔𝑔(𝑝𝑝) −𝒈𝒈(𝒑𝒑) 𝒙𝒙 𝒈𝒈(𝒑𝒑)


Legendre Transformation
• Hence Legendre transformation of a function
𝑓𝑓(𝑥𝑥) becomes :
𝑔𝑔 𝑝𝑝 = 𝑝𝑝𝑝𝑝(𝑝𝑝) − 𝑓𝑓(𝑥𝑥(𝑝𝑝))
where 𝑝𝑝 is the slope of the tangent to the
function curve.
• Summary:
• The Legendre transformation has the geometric
meaning of “the negative 𝑦𝑦 −intercept of the line
tangent to 𝑓𝑓() at the point (𝑥𝑥, 𝑓𝑓(𝑥𝑥))."
Legendre Transformation
• Since 𝑓𝑓 ′′ (𝑥𝑥) > 0 everywhere (for convex function),
there is only one tangent line for each possible slope 𝑝𝑝.
• Hence, for convex (or concave) function, the Legendre
transformation is uniquely defined.
• Each point (𝑥𝑥, 𝑓𝑓(𝑥𝑥)), has exactly one “evil twin” point
(𝑝𝑝, 𝑔𝑔(𝑝𝑝)) :
𝑥𝑥, 𝑓𝑓 𝑥𝑥 ⇔ (𝑝𝑝, 𝑔𝑔 𝑝𝑝 )
• Legendre transformations behave very badly if the
curvature of 𝑓𝑓() changes sign as 𝑥𝑥 changes, i.e.
function turns from convex to concave and vice-versa.
Legendre Transformation
• Conditions for a Legendre Transformation:
• There are two conditions for the existence of a
Legendre transformation.
• 1. The function 𝑓𝑓(𝑥𝑥) must be a well-behaved
i.e. smooth or sufficiently differentiable
function of 𝑥𝑥, defined over some part 𝐷𝐷 of the
real line.
• 2. For any 𝑥𝑥 ∈ 𝐷𝐷, 𝑓𝑓 ′′ 𝑥𝑥 > 0 (or 𝑓𝑓 ′′ 𝑥𝑥 < 0)
(i.e. everywhere in 𝐷𝐷).
Legendre Transformation
• Recipe for Legendre Transformation:
• Given a function 𝑓𝑓 𝑥𝑥 , check whether it is sufficiently
smooth and whether 𝑓𝑓 ′′ 𝑥𝑥 > 0 (or 𝑓𝑓 ′′ 𝑥𝑥 < 0 ),
within the domain of interest.
• Define a new variable 𝑝𝑝 and take 𝑝𝑝 = 𝑓𝑓 ′ (𝑥𝑥). Find 𝑥𝑥
from the relation 𝑝𝑝 = 𝑓𝑓 ′ 𝑥𝑥 i.e. find
𝑥𝑥 = 𝑥𝑥(𝑝𝑝)
from inverting 𝑝𝑝 = 𝑓𝑓 ′ 𝑥𝑥 .
• Find 𝑓𝑓(𝑥𝑥 𝑝𝑝 ) in terms of 𝑝𝑝.
• Find the Legendre transformation from:
𝑔𝑔 𝑝𝑝 = 𝑝𝑝 𝑥𝑥 𝑝𝑝 − 𝑓𝑓(𝑥𝑥 𝑝𝑝 )
Legendre Transformation
• Example-1: Find the Legendre transformation of
1
𝑓𝑓 𝑥𝑥 = 𝑥𝑥 − 𝑥𝑥0 2 and check whether any
2
information is lost.
• Solution-1:
• We have seen that, for this function, if we take
𝑝𝑝 = 𝑓𝑓 ′ 𝑥𝑥 = ℎ(𝑥𝑥) and invert this function, we get,
𝑥𝑥 = 𝑥𝑥 𝑝𝑝 = 𝑝𝑝 + 𝑥𝑥0
• Putting back into the original function, we get
1 2
1 2
𝑓𝑓 𝑥𝑥(𝑝𝑝) = 𝑝𝑝 + 𝑥𝑥0 − 𝑥𝑥0 = 𝑝𝑝
2 2
i.e. the constant 𝑥𝑥0 drops out.
Legendre Transformation
• Solution-1(contd.): Hence, functions with
different values of 𝑥𝑥0 would map to the same
1 2
final function 𝑝𝑝 , i.e. information is lost.
2
• Hence the transformation 𝑓𝑓 → ℎ−1 is not a good
transformation.
• Instead, we find the Legendre transformation
𝑓𝑓 → 𝑔𝑔, where 𝑔𝑔 is defined as:
𝑔𝑔 𝑝𝑝 = 𝑝𝑝 𝑥𝑥 𝑝𝑝 − 𝑓𝑓(𝑥𝑥 𝑝𝑝 )
• Here we have, 𝑝𝑝 = 𝑓𝑓 ′ 𝑥𝑥 = 𝑥𝑥 − 𝑥𝑥0 giving
𝑥𝑥 = 𝑥𝑥 𝑝𝑝 = 𝑝𝑝 + 𝑥𝑥0 .
Legendre Transformation
• Solution-1(contd.):
1 2
• Also,𝑓𝑓 𝑥𝑥 𝑝𝑝 = 𝑥𝑥 𝑝𝑝 − 𝑥𝑥0 = 𝑝𝑝2 /2
2
𝑝𝑝2
• 𝑔𝑔 𝑝𝑝 = 𝑝𝑝 𝑥𝑥 𝑝𝑝 − 𝑓𝑓 𝑥𝑥 𝑝𝑝 = 𝑝𝑝 𝑝𝑝 + 𝑥𝑥0 −
2
𝑝𝑝2
⇒ 𝑔𝑔 𝑝𝑝 = + 𝑝𝑝𝑥𝑥0
2
• Hence, we see that no information is lost as in the case
of ℎ−1 .
• Example-2: Find the Legendre transformation of
𝑓𝑓 𝑥𝑥 = 𝑥𝑥 ln(𝑥𝑥) for 𝑥𝑥 ∈ ℝ and 𝑥𝑥 > 0.
• Solution-2: Find 𝑝𝑝 = 𝑓𝑓 ′ 𝑥𝑥 = ln(𝑥𝑥) + 1
Legendre Transformation
• Invert the above to find 𝑥𝑥 = 𝑥𝑥(𝑝𝑝):
𝑝𝑝 = ln(𝑥𝑥) + 1 ⇒ ln(𝑥𝑥) = 𝑝𝑝 − 1 ⇒ 𝑥𝑥 = 𝑒𝑒 𝑝𝑝−1
• Find 𝑓𝑓(𝑥𝑥 𝑝𝑝 ) : 𝑓𝑓 𝑥𝑥 𝑝𝑝 = 𝑥𝑥 𝑝𝑝 ln(𝑥𝑥 𝑝𝑝 )
𝑓𝑓 𝑥𝑥 𝑝𝑝 = 𝑒𝑒 𝑝𝑝−1 ln[𝑒𝑒 𝑝𝑝−1 ] = 𝑝𝑝 − 1 𝑒𝑒 𝑝𝑝−1
• Hence, 𝑔𝑔 𝑝𝑝 = 𝑝𝑝 𝑥𝑥 𝑝𝑝 − 𝑓𝑓 𝑥𝑥 𝑝𝑝
⇒ 𝑔𝑔 𝑝𝑝 = 𝑝𝑝 𝑒𝑒 𝑝𝑝−1 − 𝑝𝑝 − 1 𝑒𝑒 𝑝𝑝−1 = 𝑒𝑒 𝑝𝑝−1
Legendre Transformation
• Properties of Legendre Transformation:
• Property-1: Slope of 𝑔𝑔(𝑝𝑝) is just 𝑥𝑥.
𝑑𝑑
• Proof: 𝑔𝑔′ 𝑝𝑝 = 𝑥𝑥 𝑝𝑝 𝑝𝑝 − 𝑓𝑓(𝑥𝑥(𝑝𝑝)) = 𝑝𝑝𝑥𝑥 ′ 𝑝𝑝 +
𝑑𝑑𝑑𝑑
𝑥𝑥 𝑝𝑝 − 𝑓𝑓 ′ 𝑥𝑥 𝑝𝑝 𝑥𝑥 ′ 𝑝𝑝 = 𝑥𝑥 𝑝𝑝
• Property-2: If 𝑓𝑓 𝑥𝑥 is convex, so also 𝑔𝑔(𝑝𝑝).
• Proof: From the chain rule and using 𝑝𝑝 = 𝑓𝑓 ′ (𝑥𝑥 𝑝𝑝 ), we
get
𝑑𝑑𝑑𝑑 𝑑𝑑 ′
1= = 𝑓𝑓 𝑥𝑥 𝑝𝑝 = 𝑓𝑓 ′′ 𝑥𝑥 𝑝𝑝 𝑥𝑥 ′ 𝑝𝑝
𝑑𝑑𝑑𝑑 𝑑𝑑𝑑𝑑
′ ′′ −1
⇒ 𝑥𝑥 𝑝𝑝 = 𝑓𝑓 𝑥𝑥 𝑝𝑝
Legendre Transformation
• Again, we have 𝑔𝑔′ 𝑝𝑝 = 𝑥𝑥(𝑝𝑝)
′′ ′ ′′ −1
• Hence, 𝑔𝑔 𝑝𝑝 = 𝑥𝑥 𝑝𝑝 = 𝑓𝑓 𝑥𝑥 𝑝𝑝 >0
since, 𝑓𝑓 was assumed to be convex.
• Hence, 𝑔𝑔(𝑝𝑝) is itself is convex in the 𝑔𝑔 − 𝑝𝑝 plane.
• Property-3: Inverse of Legendre transformation is
the same transformation.
• I.e. if we perform the Legendre transform a
second time, we recover the original function.
• Proof: Consider a function 𝑓𝑓(𝑥𝑥) whose Legendre
transformation is 𝑔𝑔 𝑝𝑝 = 𝑝𝑝𝑝𝑝 − 𝑓𝑓(𝑥𝑥)
Legendre Transformation
• Now start with the function 𝑔𝑔(𝑝𝑝) and perform its
Legendre transformation: 𝑔𝑔 𝑝𝑝 → 𝑙𝑙(𝑠𝑠)
𝑙𝑙 𝑠𝑠 = 𝑠𝑠𝑠𝑠 − 𝑔𝑔(𝑝𝑝)
where, 𝑠𝑠 = 𝑔𝑔′ (𝑝𝑝).
• This can be written as: 𝑔𝑔 𝑝𝑝 = 𝑝𝑝𝑝𝑝 − 𝑙𝑙(𝑠𝑠)
• Compare with 𝑔𝑔 𝑝𝑝 = 𝑝𝑝𝑝𝑝 − 𝑓𝑓(𝑥𝑥) and recall that,
𝑥𝑥(𝑝𝑝) = 𝑔𝑔′ (𝑝𝑝).
• Hence we can identify {𝑓𝑓, 𝑥𝑥} with {𝑙𝑙, 𝑠𝑠} (since,
𝑠𝑠 = 𝑔𝑔′ (𝑝𝑝) ).
• Hence the Legendre transformation of 𝑔𝑔(𝑝𝑝) is the
original function 𝑓𝑓(𝑥𝑥).
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Thermodynamic Potentials
• Thermodynamic systems may do work on their
environments.
• Under certain constraints, the work done on the
environment (or vice-versa) may be bounded from
above.
• This is determined by the change in an appropriately
defined quantities, called the thermodynamic
potentials.
• These are called thermodynamic potentials because
each gives a complete thermodynamic description of
the system.
• All of these have units of energy.
Thermodynamic Potentials
• Legendre Transformations of U:
• A thermodynamic system can be completely described
by its internal energy function 𝑈𝑈(𝑆𝑆, 𝑉𝑉, 𝑁𝑁).
• Here 𝑈𝑈 is internal energy, 𝑆𝑆 is entropy, 𝑉𝑉 is volume, and
𝑁𝑁 is the number of particles.
• The partial derivatives of 𝑈𝑈(𝑆𝑆, 𝑉𝑉, 𝑁𝑁) have names
suggestive of their usual physical interpretations.
• From the first law of thermodynamics, we get
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝐸𝐸𝑖𝑖𝑖𝑖𝑖𝑖 = 𝑑𝑑𝑄𝑄𝑖𝑖𝑖𝑖 + 𝑑𝑑𝑊𝑊𝑜𝑜𝑜𝑜 + 𝜇𝜇𝜇𝜇𝜇𝜇
⇒ 𝑑𝑑𝑑𝑑 = 𝑇𝑇𝑇𝑇𝑇𝑇 − 𝑃𝑃 𝑑𝑑𝑑𝑑 + 𝜇𝜇𝜇𝜇𝜇𝜇
Thermodynamic Potentials
• Here, 𝑇𝑇, 𝑃𝑃 and 𝜇𝜇 are the temperature, pressure, and
chemical potential, respectively.
• A. Transformation 𝑈𝑈, 𝑆𝑆 → {𝐴𝐴, 𝑇𝑇} : One problem with
this description is that entropomometers, even if such
things exist, are hard to find.
• Temperature-measuring devices are much more
convenient.
• Note that, the “slope” of 𝑈𝑈 = 𝑈𝑈(𝑆𝑆, 𝑉𝑉, 𝑁𝑁) with respect
to 𝑆𝑆, at constant 𝑉𝑉 and 𝑁𝑁 is just the temperature 𝑇𝑇:
𝜕𝜕𝜕𝜕
= 𝑇𝑇
𝜕𝜕𝜕𝜕 𝑉𝑉,𝑁𝑁
Thermodynamic Potentials
• Hence, we would like to use 𝑇𝑇 as a variable instead of
𝑆𝑆.
• To do this, we are tempted to use Legendre
transformation, that we learned.
• But at first we need to check whether 𝑈𝑈 = 𝑈𝑈(𝑆𝑆, 𝑉𝑉, 𝑁𝑁)
is suitable for that (i.e. either convex or concave) or
not!
𝜕𝜕𝜕𝜕 𝜕𝜕2 𝑈𝑈 𝜕𝜕𝜕𝜕
• Now, from = 𝑇𝑇 ⇒ =
𝜕𝜕𝜕𝜕 𝑉𝑉,𝑁𝑁 𝜕𝜕𝑆𝑆 2 𝑉𝑉,𝑁𝑁 𝜕𝜕𝜕𝜕 𝑉𝑉,𝑁𝑁
𝜕𝜕2 𝑈𝑈 𝜕𝜕𝜕𝜕 −1 𝜕𝜕𝜕𝜕
⇒ = > 0, since >0
𝜕𝜕𝑆𝑆 2 𝑉𝑉,𝑁𝑁 𝜕𝜕𝜕𝜕 𝑉𝑉,𝑁𝑁 𝜕𝜕𝜕𝜕 𝑉𝑉,𝑁𝑁
Thermodynamic Potentials
• The above is a physically plausible claim since:
• “Raising the temperature of a thermodynamic
system while holding everything else but 𝑈𝑈
constant will increase its entropy.”
• That is, we keep the volume and the number of
particles of a hydrostatic system (such as a gas)
constant and raise the temperature.
• The internal energy will of course increase and so
also the entropy.
• Hence, we CAN use a Legendre transformation of
𝑈𝑈(𝑆𝑆, 𝑉𝑉, 𝑁𝑁) on the variable 𝑆𝑆: i.e. 𝑈𝑈, 𝑆𝑆 → {𝐴𝐴, 𝑇𝑇}
Thermodynamic Potentials
• The other variables 𝑉𝑉, 𝑁𝑁 in 𝑈𝑈 , can also be
Legendre-transformed.
• B. Transformation 𝑈𝑈, 𝑉𝑉 → {𝐻𝐻, 𝑃𝑃}:
• Suppose we don't mind the entropy 𝑆𝑆, but 𝑉𝑉
annoys us and we prefer to use the pressure 𝑃𝑃 as
a variable in our thermodynamic potential.
• We need to check whether 𝑈𝑈(𝑆𝑆, 𝑉𝑉, 𝑁𝑁) is a
concave-up/ concave-down function for 𝑉𝑉:
𝜕𝜕𝜕𝜕
𝑃𝑃 = −
𝜕𝜕𝜕𝜕 𝑆𝑆,𝑁𝑁
Thermodynamic Potentials
𝜕𝜕 2 𝑈𝑈 𝜕𝜕𝜕𝜕
⇒ =− >0
𝜕𝜕𝑉𝑉 2 𝑆𝑆,𝑁𝑁
𝜕𝜕𝜕𝜕 𝑆𝑆,𝑁𝑁
• The above is another physically plausible claim:
• “Increasing volume decreases pressure, all other
things but 𝑈𝑈, being constant”.
• Hence, we CAN use a Legendre transformation of
𝑈𝑈(𝑆𝑆, 𝑉𝑉, 𝑁𝑁) on the variable 𝑉𝑉: i.e. 𝑈𝑈, 𝑉𝑉 → {𝐻𝐻, 𝑃𝑃}
Thermodynamic Potentials
• Helmholtz’s Free Energy 𝐴𝐴 or 𝐹𝐹:
• The Helmholtz’s free energy 𝐹𝐹 (for free), or 𝐴𝐴
(for the German word Arbeit meaning work) is
− 1 times the 𝑆𝑆 − Legendre transform of
𝑈𝑈 𝑆𝑆, 𝑉𝑉, 𝑁𝑁 : Recall: 𝑔𝑔 𝑝𝑝 = 𝑝𝑝𝑝𝑝 − 𝑓𝑓(𝑥𝑥)
𝐴𝐴 𝑇𝑇, 𝑉𝑉, 𝑁𝑁 = − 𝑇𝑇 𝑆𝑆 𝑇𝑇 − 𝑈𝑈(𝑆𝑆 𝑇𝑇 , 𝑉𝑉, 𝑁𝑁)
𝐴𝐴 = 𝐹𝐹 = 𝑈𝑈 𝑆𝑆 𝑇𝑇 , 𝑉𝑉, 𝑁𝑁 − 𝑇𝑇 𝑆𝑆(𝑇𝑇)
• Thus, in short: 𝐴𝐴 = 𝐹𝐹 = 𝑈𝑈 − 𝑇𝑇𝑇𝑇
Thermodynamic Potentials
• Enthalpy 𝐻𝐻 : The slope of 𝑈𝑈 = 𝑈𝑈(𝑆𝑆, 𝑉𝑉, 𝑁𝑁) with
respect to 𝑉𝑉 is :
𝜕𝜕𝜕𝜕
𝑃𝑃� = = −𝑃𝑃
𝜕𝜕𝜕𝜕 𝑆𝑆,𝑁𝑁
• The enthalpy 𝐻𝐻 is defined to be −1 times the
𝑉𝑉 −Legendre transform of 𝑈𝑈 𝑆𝑆, 𝑉𝑉, 𝑁𝑁 :
𝐻𝐻 𝑆𝑆, 𝑃𝑃�, 𝑁𝑁 = − 𝑃𝑃�𝑉𝑉 𝑃𝑃� − 𝑈𝑈 𝑆𝑆, 𝑉𝑉 𝑃𝑃� , 𝑁𝑁
⇒ 𝐻𝐻 𝑆𝑆, 𝑃𝑃, 𝑁𝑁 = 𝑈𝑈 𝑆𝑆, 𝑉𝑉 𝑃𝑃 , 𝑁𝑁 + 𝑃𝑃𝑃𝑃(𝑃𝑃)
Thermodynamic Potentials
• The + before 𝑃𝑃𝑃𝑃(𝑃𝑃) is a side effect of defining 𝑃𝑃
with the wrong sign!
• In short: 𝐻𝐻 = 𝑈𝑈 + 𝑃𝑃𝑃𝑃
• 𝑃𝑃� = 𝜕𝜕𝜕𝜕/𝜕𝜕𝜕𝜕 𝑆𝑆,𝑁𝑁 is the negative of pressure and
may be called “vacuosity”.
• Gibb’s Free Energy 𝐺𝐺 :
• If we can 𝑆𝑆 − Legendre transform 𝑈𝑈 and 𝑉𝑉 −
Legendre transform 𝑈𝑈, can we do both?
• The answer is YES!
Thermodynamic Potentials
• We first 𝑆𝑆 −Legendre transform 𝑈𝑈 𝑆𝑆, 𝑉𝑉, 𝑁𝑁 to get 𝐴𝐴
(i.e. 𝐹𝐹 ) and then 𝑉𝑉 − Legendre transform the
𝑈𝑈 = 𝑈𝑈(𝑆𝑆 𝑇𝑇 , 𝑉𝑉, 𝑁𝑁) inside 𝐴𝐴 and put a −1 sign in front ,
at each step, to get the Gibb’s free energy, 𝐺𝐺:
𝑈𝑈 𝑆𝑆, 𝑉𝑉, 𝑁𝑁 → 𝑈𝑈 𝑆𝑆 𝑇𝑇 , 𝑉𝑉, 𝑁𝑁 − 𝑇𝑇 𝑆𝑆 𝑇𝑇
[-1 times 𝑆𝑆 −Legendre transform ]
→ 𝑈𝑈 𝑆𝑆 𝑇𝑇 , 𝑉𝑉 𝑃𝑃 , 𝑁𝑁 + 𝑃𝑃𝑃𝑃 𝑃𝑃 − 𝑇𝑇 𝑆𝑆 𝑇𝑇
[-1 times 𝑉𝑉 −Legendre transform of 𝑈𝑈 𝑆𝑆 𝑇𝑇 , 𝑉𝑉, 𝑁𝑁 ]
• Hence
𝐺𝐺 𝑇𝑇, 𝑃𝑃, 𝑁𝑁 = 𝑈𝑈 𝑆𝑆 𝑇𝑇 , 𝑉𝑉 𝑃𝑃 , 𝑁𝑁 + 𝑃𝑃𝑃𝑃 𝑃𝑃 − 𝑇𝑇 𝑆𝑆 𝑇𝑇
• In short, 𝐺𝐺 = 𝑈𝑈 − 𝑇𝑇𝑇𝑇 + 𝑃𝑃𝑃𝑃
Thermodynamic Potentials
• Landau Potential or Grand Canonical Potential 𝚽𝚽 :
• We can similarly perform an 𝑁𝑁 − Legendre
transformation of 𝑈𝑈(𝑆𝑆, 𝑉𝑉, 𝑁𝑁) at the end.
• Starting from 𝑈𝑈(𝑆𝑆, 𝑉𝑉, 𝑁𝑁), we (-1 times) 𝑆𝑆 −Legendre
transform 𝑈𝑈(𝑆𝑆, 𝑉𝑉, 𝑁𝑁) to get:
𝑈𝑈 𝑆𝑆, 𝑉𝑉, 𝑁𝑁 → 𝑈𝑈 𝑆𝑆 𝑇𝑇 , 𝑉𝑉, 𝑁𝑁 − 𝑇𝑇 𝑆𝑆 𝑇𝑇
• Next, we (-1 times) 𝑉𝑉 − Legendre transform
𝑈𝑈(𝑆𝑆(𝑇𝑇), 𝑉𝑉, 𝑁𝑁) inside the above expression to get:
𝑈𝑈 𝑆𝑆 𝑇𝑇 , 𝑉𝑉, 𝑁𝑁 − 𝑇𝑇 𝑆𝑆 𝑇𝑇
→ 𝑈𝑈 𝑆𝑆 𝑇𝑇 , 𝑉𝑉 𝑃𝑃 , 𝑁𝑁 + 𝑃𝑃𝑃𝑃 𝑃𝑃 − 𝑇𝑇 𝑆𝑆 𝑇𝑇
Thermodynamic Potentials
• Next we (-1 times) 𝑁𝑁 − Legendre transform
𝑈𝑈(𝑆𝑆(𝑇𝑇), 𝑉𝑉(𝑃𝑃), 𝑁𝑁) inside the above expression to
get:
𝑈𝑈 𝑆𝑆 𝑇𝑇 , 𝑉𝑉 𝑃𝑃 , 𝑁𝑁 + 𝑃𝑃𝑃𝑃 𝑃𝑃 − 𝑇𝑇 𝑆𝑆 𝑇𝑇
→ 𝑈𝑈 𝑆𝑆 𝑇𝑇 , 𝑉𝑉 𝑃𝑃 , 𝑁𝑁(𝜇𝜇) − 𝜇𝜇𝜇𝜇 + 𝑃𝑃𝑃𝑃 𝑃𝑃 − 𝑇𝑇 𝑆𝑆 𝑇𝑇
• At the end we get:
Φ 𝑇𝑇, 𝑃𝑃, 𝜇𝜇
= 𝑈𝑈 𝑆𝑆 𝑇𝑇 , 𝑉𝑉 𝑃𝑃 , 𝑁𝑁(𝜇𝜇) − 𝑇𝑇𝑇𝑇 + 𝑃𝑃𝑃𝑃 − 𝜇𝜇𝜇𝜇
• In short: Φ = 𝑈𝑈 − 𝑇𝑇𝑇𝑇 + 𝑃𝑃𝑃𝑃 − 𝜇𝜇𝜇𝜇
Thermodynamic Potentials
• Hence, 𝑑𝑑Φ = 𝑑𝑑𝑑𝑑 − 𝑇𝑇𝑇𝑇𝑇𝑇 − 𝑆𝑆𝑆𝑆𝑆𝑆 + 𝑃𝑃𝑃𝑃𝑃𝑃 +
𝑉𝑉𝑉𝑉𝑉𝑉 − 𝜇𝜇𝜇𝜇𝜇𝜇 + 𝑁𝑁 𝑑𝑑𝑑𝑑
• Using the expression for 𝑑𝑑𝑑𝑑 from the first law,
we get:
𝑑𝑑Φ = 𝑇𝑇𝑇𝑇𝑇𝑇 − 𝑃𝑃𝑃𝑃𝑃𝑃 + 𝜇𝜇𝜇𝜇𝜇𝜇 − 𝑇𝑇𝑇𝑇𝑇𝑇 − 𝑆𝑆𝑆𝑆𝑆𝑆
+ 𝑃𝑃𝑃𝑃𝑃𝑃 + 𝑉𝑉𝑉𝑉𝑉𝑉 − 𝜇𝜇𝜇𝜇𝜇𝜇 + 𝑁𝑁 𝑑𝑑𝑑𝑑
𝑑𝑑Φ = −𝑆𝑆𝑆𝑆𝑆𝑆 + 𝑉𝑉𝑉𝑉𝑉𝑉 + 𝑁𝑁 𝑑𝑑𝑑𝑑
hence the dependence of Φ on 𝑇𝑇, 𝑃𝑃, 𝜇𝜇.
Thermodynamic Potentials
• Properties of Thermodynamic Potentials:
• 1. The different thermodynamic potentials are
equivalent in the sense that no information is lost in
the transformation (i.e. Legendre transformation) from
one potential to the other.
• 2. All the thermodynamic potentials have the
dimensions of the energy.
• 3. The thermodynamic potentials are expressed in
thermodynamic coordinates amenable to the
experimental situation at hand.
• 4. All the thermodynamic properties of a system can be
calculated by differentiation of these potentials only.
Thermodynamic Potentials
• 5. The choice of 𝑈𝑈 , 𝐻𝐻 , 𝐴𝐴 , and 𝐺𝐺 as the
fundamental set of functions has the advantage
that all four functions, are conserved.
• Historical Aside:
• This remarkable formalism and procedure was
introduced into thermodynamics during the
1870s by J. Willard Gibbs, Professor of
Mathematical Physics at Yale during his entire
career.
• But the names of the functions and their symbols
were chosen by other scientists.
Thermodynamic Potentials
• Equivalent forms of the First Law: In terms of the
thermodynamics potentials, we can write the first law in
four equivalent forms:
• 𝑈𝑈 = 𝑈𝑈(𝑆𝑆, 𝑉𝑉, 𝑁𝑁): 𝑑𝑑𝑑𝑑 = 𝑇𝑇𝑇𝑇𝑇𝑇 − 𝑃𝑃𝑃𝑃𝑃𝑃 + 𝜇𝜇𝜇𝜇𝜇𝜇
• 𝐴𝐴 = 𝐴𝐴 𝑇𝑇, 𝑉𝑉, 𝑁𝑁 = 𝑈𝑈 − 𝑇𝑇𝑇𝑇:
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 = 𝑑𝑑 𝑈𝑈 − 𝑇𝑇𝑇𝑇 = 𝑑𝑑𝑑𝑑 − 𝑇𝑇𝑇𝑇𝑇𝑇 − 𝑆𝑆𝑆𝑆𝑆𝑆
⇒ 𝑑𝑑𝑑𝑑 = −𝑆𝑆𝑆𝑆𝑆𝑆 − 𝑃𝑃𝑃𝑃𝑃𝑃 + 𝜇𝜇𝜇𝜇𝜇𝜇
• 𝐻𝐻 = 𝐻𝐻 𝑆𝑆, 𝑃𝑃, 𝑁𝑁 = 𝑈𝑈 + 𝑃𝑃𝑃𝑃:
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑃𝑃𝑃𝑃𝑃𝑃 + 𝑉𝑉𝑉𝑉𝑉𝑉 = 𝑇𝑇𝑇𝑇𝑇𝑇 + 𝑉𝑉𝑉𝑉𝑉𝑉 + 𝜇𝜇𝜇𝜇𝜇𝜇
• 𝐺𝐺 = 𝐺𝐺 𝑇𝑇, 𝑃𝑃, 𝑁𝑁 = 𝑈𝑈 − 𝑇𝑇𝑇𝑇 + 𝑃𝑃𝑃𝑃:
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 − 𝑇𝑇𝑇𝑇𝑇𝑇 − 𝑆𝑆𝑆𝑆𝑆𝑆 + 𝑃𝑃𝑃𝑃𝑃𝑃 + 𝑉𝑉𝑉𝑉𝑉𝑉
⇒ 𝑑𝑑𝑑𝑑 = −𝑆𝑆𝑆𝑆𝑆𝑆 + 𝑉𝑉𝑉𝑉𝑉𝑉 + 𝜇𝜇𝜇𝜇𝜇𝜇
Maxwell’s Relations
• Maxwell relations are conditions equating certain
derivatives of the state variables which follow from the
exactness of the differentials of the various state functions.
• Relations from 𝑈𝑈(𝑆𝑆, 𝑉𝑉, 𝑁𝑁):
• The internal energy is a state function and can be expressed
as 𝑈𝑈 = 𝑈𝑈(𝑆𝑆, 𝑉𝑉, 𝑁𝑁)
• From the first law: 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑄𝑄𝑖𝑖𝑖𝑖 + 𝑑𝑑𝑊𝑊𝑜𝑜𝑜𝑜 + 𝜇𝜇𝜇𝜇𝜇𝜇
⇒ 𝑑𝑑𝑑𝑑 = 𝑇𝑇𝑇𝑇𝑇𝑇 − 𝑃𝑃𝑃𝑃𝑃𝑃 + 𝜇𝜇𝜇𝜇𝜇𝜇
• Therefore, we have
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑇𝑇 = ; −𝑃𝑃 = ; 𝜇𝜇 =
𝜕𝜕𝜕𝜕 𝑉𝑉,𝑁𝑁 𝜕𝜕𝜕𝜕 𝑆𝑆,𝑁𝑁 𝜕𝜕𝜕𝜕 𝑆𝑆,𝑉𝑉
Maxwell’s Relations
• Taking the mixed second derivative we find,
𝜕𝜕 2 𝑈𝑈 𝜕𝜕 2 𝑈𝑈 𝜕𝜕𝑇𝑇 𝜕𝜕𝑃𝑃
= = =−
𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕 𝜕𝜕𝑉𝑉𝜕𝜕𝑆𝑆 𝜕𝜕𝜕𝜕 𝑆𝑆,𝑁𝑁 𝜕𝜕𝜕𝜕 𝑉𝑉,𝑁𝑁
𝜕𝜕 2 𝑈𝑈 𝜕𝜕 2 𝑈𝑈 𝜕𝜕𝜕𝜕 𝜕𝜕𝜇𝜇
= = =
𝜕𝜕𝜕𝜕𝜕𝜕𝑁𝑁 𝜕𝜕𝑁𝑁𝜕𝜕𝜕𝜕 𝜕𝜕𝑁𝑁 𝑆𝑆,𝑉𝑉
𝜕𝜕𝜕𝜕 𝑉𝑉,𝑁𝑁
𝜕𝜕 2 𝑈𝑈 𝜕𝜕 2 𝑈𝑈 𝜕𝜕𝑃𝑃 𝜕𝜕𝜕𝜕
= =− =
𝜕𝜕𝑉𝑉𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕𝜕𝜕𝑉𝑉 𝜕𝜕𝜕𝜕 𝑆𝑆,𝑉𝑉
𝜕𝜕𝑉𝑉 𝑆𝑆,𝑁𝑁
Maxwell’s Relations
• Relations from 𝐹𝐹(𝑇𝑇, 𝑉𝑉, 𝑁𝑁) or 𝐴𝐴(𝑇𝑇, 𝑉𝑉, 𝑁𝑁) :
• The Helmholtz free energy 𝐹𝐹 or 𝐴𝐴 is a state
function whose change is given by the first
law: 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 = −𝑆𝑆𝑆𝑆𝑆𝑆 − 𝑃𝑃𝑃𝑃𝑃𝑃 + 𝜇𝜇𝜇𝜇𝜇𝜇
• Therefore, we have
𝜕𝜕𝐹𝐹 𝜕𝜕𝐹𝐹 𝜕𝜕𝐹𝐹
𝑆𝑆 = − ; −𝑃𝑃 = ; 𝜇𝜇 =
𝜕𝜕𝑇𝑇 𝑉𝑉,𝑁𝑁 𝜕𝜕𝜕𝜕 𝑇𝑇,𝑁𝑁 𝜕𝜕𝜕𝜕 𝑇𝑇,𝑉𝑉
Maxwell’s Relations
• Taking the mixed second derivative we find,
𝜕𝜕 2 𝐹𝐹 𝜕𝜕 2 𝐹𝐹 𝜕𝜕𝑆𝑆 𝜕𝜕𝜕𝜕
= =− =−
𝜕𝜕𝑇𝑇𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕𝜕𝜕𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇,𝑁𝑁 𝜕𝜕𝑇𝑇 𝑉𝑉,𝑁𝑁
𝜕𝜕 2 𝐹𝐹 𝜕𝜕 2 𝐹𝐹 𝜕𝜕𝑆𝑆 𝜕𝜕𝜕𝜕
= =− =
𝜕𝜕𝑇𝑇𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕𝜕𝜕𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇,𝑉𝑉
𝜕𝜕𝑇𝑇 𝑉𝑉,𝑁𝑁
𝜕𝜕 2 𝐹𝐹 𝜕𝜕 2 𝐹𝐹 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
= =− =
𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑇𝑇,𝑉𝑉
𝜕𝜕𝜕𝜕 𝑇𝑇,𝑁𝑁
Maxwell’s Relations
• Relations from 𝐻𝐻(𝑆𝑆, 𝑃𝑃, 𝑁𝑁):
• The enthalpy is a state function whose change
is given by the first law:𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑃𝑃𝑃𝑃𝑃𝑃 +
𝑉𝑉𝑉𝑉𝑉𝑉 = 𝑇𝑇𝑇𝑇𝑇𝑇 + 𝑉𝑉𝑉𝑉𝑉𝑉 + 𝜇𝜇𝜇𝜇𝜇𝜇
• Therefore, we have
𝜕𝜕𝐻𝐻 𝜕𝜕𝐻𝐻 𝜕𝜕𝐻𝐻
𝑇𝑇 = ; 𝑉𝑉 = ; 𝜇𝜇 =
𝜕𝜕𝑆𝑆 𝑃𝑃,𝑁𝑁 𝜕𝜕𝑃𝑃 𝑆𝑆,𝑁𝑁 𝜕𝜕𝜕𝜕 𝑆𝑆,𝑃𝑃
Maxwell’s Relations
• Taking the mixed second derivative we find,
𝜕𝜕 2 𝐻𝐻 𝜕𝜕 2 𝐻𝐻 𝜕𝜕𝑇𝑇 𝜕𝜕𝑉𝑉
= = =
𝜕𝜕𝑆𝑆𝜕𝜕𝑃𝑃 𝜕𝜕𝑃𝑃𝜕𝜕𝑆𝑆 𝜕𝜕𝑃𝑃 𝑆𝑆,𝑁𝑁 𝜕𝜕𝑆𝑆 𝑃𝑃,𝑁𝑁
𝜕𝜕 2 𝐻𝐻 𝜕𝜕 2 𝐻𝐻 𝜕𝜕𝑇𝑇 𝜕𝜕𝜕𝜕
= = =
𝜕𝜕𝑆𝑆𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕𝜕𝜕𝑆𝑆 𝜕𝜕𝜕𝜕 𝑆𝑆,𝑃𝑃
𝜕𝜕𝑆𝑆 𝑃𝑃,𝑁𝑁
𝜕𝜕 2 𝐻𝐻 𝜕𝜕 2 𝐻𝐻 𝜕𝜕𝑉𝑉 𝜕𝜕𝜕𝜕
= = =
𝜕𝜕𝑃𝑃𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕𝜕𝜕𝑃𝑃 𝜕𝜕𝜕𝜕 𝑆𝑆,𝑃𝑃
𝜕𝜕𝑃𝑃 𝑆𝑆,𝑁𝑁
Maxwell’s Relations
• Relations from 𝐺𝐺(𝑇𝑇, 𝑃𝑃, 𝑁𝑁):
• The Gibb’s free energy is a state function
whose change is given by the first law:
𝑑𝑑𝑑𝑑 = −𝑆𝑆𝑆𝑆𝑆𝑆 + 𝑉𝑉𝑉𝑉𝑉𝑉 + 𝜇𝜇𝜇𝜇𝜇𝜇
• Therefore, we have
𝜕𝜕𝐺𝐺 𝜕𝜕𝐺𝐺 𝜕𝜕𝐺𝐺
−𝑆𝑆 = ; 𝑉𝑉 = ; 𝜇𝜇 =
𝜕𝜕𝑇𝑇 𝑃𝑃,𝑁𝑁 𝜕𝜕𝜕𝜕 𝑇𝑇,𝑁𝑁 𝜕𝜕𝜕𝜕 𝑇𝑇,𝑃𝑃
Maxwell’s Relations
• Taking the mixed second derivative we find,
𝜕𝜕 2 𝐺𝐺 𝜕𝜕 2 𝐺𝐺 𝜕𝜕𝑆𝑆 𝜕𝜕𝜕𝜕
= =− =
𝜕𝜕𝑇𝑇𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕𝜕𝜕𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇,𝑁𝑁 𝜕𝜕𝑇𝑇 𝑃𝑃,𝑁𝑁
𝜕𝜕 2 𝐺𝐺 𝜕𝜕 2 𝐺𝐺 𝜕𝜕𝑆𝑆 𝜕𝜕𝜕𝜕
= =− =
𝜕𝜕𝑇𝑇𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕𝜕𝜕𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇,𝑃𝑃
𝜕𝜕𝑇𝑇 𝑃𝑃,𝑁𝑁
𝜕𝜕 2 𝐺𝐺 𝜕𝜕 2 𝐺𝐺 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
= = =
𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑇𝑇,𝑃𝑃
𝜕𝜕𝜕𝜕 𝑇𝑇,𝑁𝑁
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Thermodynamic Potentials
• We have, the familiar thermodynamic potentials:
𝑈𝑈 = 𝑈𝑈 𝑆𝑆, 𝑉𝑉, 𝑁𝑁
𝐴𝐴 = 𝐹𝐹 = 𝐴𝐴 𝑇𝑇, 𝑉𝑉, 𝑁𝑁 = 𝑈𝑈 − 𝑇𝑇𝑇𝑇
𝐻𝐻 = 𝐻𝐻 𝑆𝑆, 𝑃𝑃, 𝑁𝑁 = 𝑈𝑈 + 𝑃𝑃𝑃𝑃
𝐺𝐺 = 𝐺𝐺 𝑇𝑇, 𝑃𝑃, 𝑁𝑁 = 𝑈𝑈 − 𝑇𝑇𝑇𝑇 + 𝑃𝑃𝑃𝑃
Φ = Φ 𝑇𝑇, 𝑃𝑃, 𝜇𝜇 = 𝑈𝑈 − 𝑇𝑇𝑇𝑇 + 𝑃𝑃𝑃𝑃 − 𝜇𝜇𝜇𝜇
• Together with the five equivalent forms of the first law:
𝑑𝑑𝑑𝑑 = 𝑇𝑇𝑇𝑇𝑇𝑇 − 𝑃𝑃𝑃𝑃𝑃𝑃 + 𝜇𝜇𝜇𝜇𝜇𝜇
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 = −𝑆𝑆𝑆𝑆𝑆𝑆 − 𝑃𝑃𝑃𝑃𝑃𝑃 + 𝜇𝜇𝜇𝜇𝜇𝜇
𝑑𝑑𝑑𝑑 = 𝑇𝑇𝑇𝑇𝑇𝑇 + 𝑉𝑉𝑉𝑉𝑉𝑉 + 𝜇𝜇𝜇𝜇𝜇𝜇
𝑑𝑑𝑑𝑑 = −𝑆𝑆𝑆𝑆𝑆𝑆 + 𝑉𝑉𝑉𝑉𝑉𝑉 + 𝜇𝜇𝜇𝜇𝜇𝜇
𝑑𝑑Φ = −𝑆𝑆𝑆𝑆𝑆𝑆 + 𝑉𝑉𝑉𝑉𝑉𝑉 + 𝑁𝑁𝑁𝑁𝑁𝑁
Thermodynamic Potentials
• Helmholtz Free Energy- Physical Interpretation:
• Helmholtz function 𝐴𝐴 = 𝐹𝐹 = 𝑈𝑈 − 𝑇𝑇𝑇𝑇 is a useful
function in isothermal processes.
• A: Suppose we spontaneously create a system by
bringing together the constituents while keeping
it in constant contact with a thermal reservoir at
temperature 𝑇𝑇.
• As we bring in the constituents (e.g. particle) of
the system, they will absorb heat from the
thermal reservoir and start thermal motion.
Thermodynamic Potentials
• Thus in the process of creation of the system,
heat absorbed is 𝑄𝑄 = ∫ 𝑑𝑑𝑑𝑑 = ∫ 𝑇𝑇𝑇𝑇𝑇𝑇 = 𝑇𝑇 ∫ 𝑑𝑑𝑑𝑑 =
𝑇𝑇𝑇𝑇
• Therefore, we don’t have to supply the full
internal energy 𝑈𝑈, but rather only 𝑈𝑈 − 𝑄𝑄, since
the system receives heat energy 𝑄𝑄 = 𝑇𝑇𝑇𝑇 from the
reservoir.
• Thus, we must perform work 𝑊𝑊𝑜𝑜𝑜𝑜 = 𝑈𝑈 − 𝑇𝑇𝑇𝑇 to
create our system, if it is constantly in equilibrium
with a reservoir at temperature 𝑇𝑇.
Thermodynamic Potentials
• The quantity 𝑈𝑈 − 𝑇𝑇𝑇𝑇 is known as the Helmholtz
free energy, 𝐹𝐹 or 𝐴𝐴.
• B: Consider two systems interacting mechanically
and diffusively but held at constant temperature.
• For each, the change in Helmholtz function is
given by: (from 𝑑𝑑𝑑𝑑 = 0 in 𝑑𝑑𝑑𝑑 = −𝑆𝑆𝑆𝑆𝑆𝑆 − 𝑃𝑃𝑃𝑃𝑃𝑃 +
𝜇𝜇𝜇𝜇𝜇𝜇)
𝑑𝑑𝐹𝐹𝑖𝑖 = −𝑃𝑃𝑖𝑖 𝑑𝑑𝑉𝑉𝑖𝑖 + 𝜇𝜇𝑖𝑖 𝑑𝑑𝑁𝑁𝑖𝑖
where the subscript 𝑖𝑖 denotes the number/index
of the system, i.e. 𝑖𝑖 = 1,2.
Thermodynamic Potentials
• Since the volume and particles gained by one are
lost by the other, we have:
𝑑𝑑𝑉𝑉1 = −𝑑𝑑𝑉𝑉2 and 𝑑𝑑𝑁𝑁1 = −𝑑𝑑𝑁𝑁2
• The total change in 𝐹𝐹 is: 𝑑𝑑𝐹𝐹𝑡𝑡𝑡𝑡𝑡𝑡 = 𝑑𝑑𝐹𝐹1 + 𝑑𝑑𝐹𝐹2
⇒ 𝑑𝑑𝐹𝐹𝑡𝑡𝑡𝑡𝑡𝑡 = − 𝑃𝑃1 − 𝑃𝑃2 𝑑𝑑𝑉𝑉1 + 𝜇𝜇1 − 𝜇𝜇2 𝑑𝑑𝑁𝑁1
• The volume of the system having greater pressure
will increase.
• The particles flow towards
the system having smaller
chemical potential.
𝐴𝐴2 ≡ system -2
𝐴𝐴1 ≡ system -1
Thermodynamic Potentials
• Thus, if 𝑃𝑃1 > 𝑃𝑃2 , 𝑑𝑑𝑉𝑉1 > 0 i.e. the system with greater
pressure (here system 1) gains volume.
⇒ − 𝑃𝑃1 − 𝑃𝑃2 𝑑𝑑𝑉𝑉1 < 0
• Again, if 𝜇𝜇1 > 𝜇𝜇2 , then 𝑑𝑑𝑁𝑁1 < 0, i.e. 𝑁𝑁1 decreases.
⇒ 𝜇𝜇1 − 𝜇𝜇2 𝑑𝑑𝑁𝑁1 < 0
⇒ 𝑑𝑑𝐹𝐹𝑡𝑡𝑡𝑡𝑡𝑡 = − 𝑃𝑃1 − 𝑃𝑃2 𝑑𝑑𝑉𝑉1 + 𝜇𝜇1 − 𝜇𝜇2 𝑑𝑑𝑁𝑁1 < 0
• That is, for systems interacting isothermally and
mechanically and diffusively, changes in the total
Helmholtz function must be negative.
• The Helmholtz function reaches a minimum at
equilibrium.
Thermodynamic Potentials
• That the Helmholtz function must reach a minimum is
a consequence of the second law of thermodynamics.
• B: For a reversible isothermal and non-diffusive
process (i.e. 𝑑𝑑𝑑𝑑 = 0 , 𝑑𝑑𝑑𝑑 = 0 in 𝑑𝑑𝑑𝑑 = −𝑆𝑆𝑆𝑆𝑆𝑆 − 𝑃𝑃𝑃𝑃𝑃𝑃 +
𝜇𝜇𝜇𝜇𝜇𝜇)
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 = −𝑃𝑃𝑃𝑃𝑃𝑃
𝑓𝑓
⇒ 𝐴𝐴𝑓𝑓 − 𝐴𝐴𝑖𝑖 = − � 𝑃𝑃𝑃𝑃𝑃𝑃 𝑇𝑇,𝑁𝑁 = 𝑊𝑊𝑜𝑜𝑜𝑜 �
𝑇𝑇,𝑁𝑁 𝑇𝑇,𝑁𝑁
𝑖𝑖
• The increase of the Helmholtz function during a
reversible isothermal non-diffusive process equals the
work done on the system.
Thermodynamic Potentials
• For any finite reversible isothermal non-
diffusive process, we may write:
Δ𝐴𝐴� = Δ𝑈𝑈� − 𝑇𝑇 Δ𝑆𝑆�
𝑇𝑇,𝑁𝑁 𝑇𝑇,𝑁𝑁 𝑇𝑇,𝑁𝑁
= Δ𝑈𝑈� − Δ𝑄𝑄� = Δ𝑊𝑊𝑜𝑜𝑜𝑜 �
𝑇𝑇,𝑁𝑁 𝑇𝑇,𝑁𝑁 𝑇𝑇,𝑁𝑁
• The increase of the Helmholtz energy Δ𝐴𝐴| 𝑇𝑇,𝑁𝑁
of a system equals the amount of isothermal
work Δ𝑊𝑊𝑜𝑜𝑜𝑜 | 𝑇𝑇,𝑁𝑁 that is performed on the
system in a reversible process.
Thermodynamic Potentials
• Equivalently, the decrease of the Helmholtz
energy Δ𝐴𝐴| 𝑇𝑇,𝑁𝑁 < 0 of a system equals the
amount of isothermal work Δ𝑊𝑊𝑏𝑏𝑏𝑏 � < 0 that is
𝑇𝑇,𝑁𝑁
performed by the system in a reversible process.
• If a system does isothermal work (i.e. work done
by the system), the internal energy 𝑈𝑈𝑇𝑇,𝑁𝑁 also
decreases.
• But the decrease Δ𝑈𝑈𝑇𝑇,𝑁𝑁 does not always equal
the work, that the system can perform.
• If Δ𝑄𝑄𝑇𝑇,𝑁𝑁 = 0 then Δ𝑊𝑊𝑇𝑇,𝑁𝑁 = Δ𝑈𝑈𝑇𝑇,𝑁𝑁
Thermodynamic Potentials
• If, Δ𝑄𝑄𝑇𝑇,𝑁𝑁 > 0 ⇒ Δ𝑊𝑊𝑇𝑇,𝑁𝑁 > Δ𝑈𝑈𝑇𝑇,𝑁𝑁
• If, Δ𝑄𝑄𝑇𝑇,𝑁𝑁 < 0 ⇒ Δ𝑊𝑊𝑇𝑇,𝑁𝑁 < Δ𝑈𝑈𝑇𝑇,𝑁𝑁
• Since, 𝐴𝐴 or 𝐹𝐹 represents the work done in
non-diffusive, isothermal process, it is called
the Helmholtz FREE energy.
• C: For a reversible isothermal, isochoric and
non-diffusive process,
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 = −𝑆𝑆𝑆𝑆𝑆𝑆 − 𝑃𝑃𝑃𝑃𝑃𝑃 + 𝜇𝜇𝜇𝜇𝜇𝜇 = 0
Thermodynamic Potentials
• ⇒ 𝐴𝐴𝑖𝑖 − 𝐴𝐴𝑓𝑓 = 0 or 𝐴𝐴 = const.
𝑇𝑇,𝑁𝑁,𝑉𝑉
• D. Again, from 𝑑𝑑𝐴𝐴 = 𝑑𝑑𝑑𝑑 = −𝑆𝑆𝑆𝑆𝑆𝑆 − 𝑃𝑃𝑃𝑃𝑃𝑃 +
𝜇𝜇𝜇𝜇𝜇𝜇, we have:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
= −𝑃𝑃 and = −𝑆𝑆
𝜕𝜕𝜕𝜕 𝑇𝑇,𝑁𝑁 𝜕𝜕𝜕𝜕 𝑉𝑉,𝑁𝑁

• Thus, the pressure and entropy are found by


direct differentiation of the Helmholtz free
energy 𝐴𝐴.
Thermodynamic Potentials
• Summary: Helmholtz free energy 𝑨𝑨 = 𝑭𝑭 = 𝑼𝑼 − 𝑻𝑻𝑻𝑻
• 1. 𝐴𝐴 = 𝐹𝐹 = 𝑈𝑈 − 𝑇𝑇𝑇𝑇 equals to the work done in
creating a system while keeping it in constant contact
with a thermal reservoir at temperature 𝑇𝑇.
• 2. The increase of the Helmholtz function during a
reversible isothermal non-diffusive process equals the
work done on the system.
• 3. For systems interacting isothermally and
mechanically and diffusively, changes in the total
Helmholtz free energy must be negative.
• 4. Helmholtz function reaches a minimum at
equilibrium.
Thermodynamic Potentials
• 5. For a reversible isothermal, isochoric and non-
diffusive process the Helmholtz free energy is a
constant.
• 6. Pressure and entropy are found by direct
differentiation of the Helmholtz free energy 𝐴𝐴.
• Gibbs Free Energy- Physical Interpretation:
• Gibbs free energy is relevant in a wide variety of
diffusive processes that reach equilibrium under
isothermal and isobaric constraints i.e. for
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 = 0.
Thermodynamic Potentials
• A. Suppose we create a thermodynamic system
by bringing in the constituents at conditions of
constant temperature 𝑇𝑇 and constant pressure 𝑃𝑃.
• For example, we may create the system inside a
chamber with a movable piston that is exerting a
constant pressure 𝑃𝑃 and in contact with a
thermal reservoir.
• The constituents of the system (e.g. particles) will
absorb heat from the thermal reservoir at 𝑇𝑇.
Thermodynamic Potentials
• The heat absorbed is 𝑄𝑄 = ∫ 𝑑𝑑𝑑𝑑 = ∫ 𝑇𝑇𝑇𝑇𝑇𝑇 = 𝑇𝑇 ∫ 𝑑𝑑𝑑𝑑 =
𝑇𝑇𝑇𝑇.
• Furthermore, while creating the system, besides
supplying internal energy, we must also perform work
∫ 𝑃𝑃𝑃𝑃𝑃𝑃 = 𝑃𝑃 ∫ 𝑑𝑑𝑑𝑑 = 𝑃𝑃𝑃𝑃 in order to make room for it.
• This extra work has to be done because the system is
being created against an applied constant pressure.
• Had we created the system in vacuum (with no applied
pressure on it), we would have needed only the
potential and kinetic energies of the particles, as parts
of the internal energy.
Thermodynamic Potentials
• In other words, we must perform a work
𝑊𝑊 = 𝑈𝑈 + 𝑃𝑃𝑃𝑃 − 𝑇𝑇𝑇𝑇 in assembling our system,
if it is constantly in equilibrium with a
reservoir at temperature 𝑇𝑇 and at constant
pressure 𝑃𝑃.
• The quantity 𝑈𝑈 − 𝑇𝑇𝑇𝑇 + 𝑃𝑃𝑃𝑃 is known as the
Gibbs free energy.
Thermodynamic Potentials
• B: Consider two systems interacting diffusively.
The change in Gibbs free energy for either system
is 𝑑𝑑𝑑𝑑 = 𝜇𝜇𝜇𝜇𝜇𝜇 (from 𝑑𝑑𝑑𝑑 = −𝑆𝑆𝑆𝑆𝑆𝑆 + 𝑉𝑉𝑉𝑉𝑉𝑉 + 𝜇𝜇𝜇𝜇𝜇𝜇,
with 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 = 0)
• The particles gained by one are lost by the other
i.e. 𝑑𝑑𝑁𝑁1 = −𝑑𝑑𝑁𝑁2
• Hence, the change in Gibbs free energy for the
combined system is:
𝑑𝑑𝐺𝐺𝑡𝑡𝑡𝑡𝑡𝑡 = 𝑑𝑑𝐺𝐺1 + 𝑑𝑑𝐺𝐺2 = 𝜇𝜇1 − 𝜇𝜇2 𝑑𝑑𝑁𝑁1
Thermodynamic Potentials
• But particles are gained by the system having the lower
chemical potential.
• Hence, if 𝜇𝜇1 > 𝜇𝜇2 , 𝑑𝑑𝑁𝑁1 < 0 which implies:
𝑑𝑑𝐺𝐺𝑡𝑡𝑡𝑡𝑡𝑡 = 𝜇𝜇1 − 𝜇𝜇2 𝑑𝑑𝑁𝑁1 < 0
• Thus for systems interacting isothermally, isobarically
and diffusively, changes in the total Gibbs free energy
must be negative.
• Hence, the combined system will reach an equilibrium,
when its Gibbs free energy will reach a minimum.
• When a system is in diffusive equilibrium its Gibbs
free energy is a minimum for that particular
temperature and pressure.
Thermodynamic Potentials
• C. For a reversible isothermal, isobaric and non-
diffusive process,
𝑑𝑑𝐺𝐺 = −𝑆𝑆𝑆𝑆𝑆𝑆 + 𝑉𝑉𝑉𝑉𝑉𝑉 + 𝜇𝜇𝜇𝜇𝜇𝜇 = 0
• ⇒ 𝐺𝐺𝑖𝑖 − 𝐺𝐺𝑓𝑓 = 0 or 𝐺𝐺 = const.
𝑇𝑇,𝑃𝑃,𝑁𝑁
• E. Again, from 𝑑𝑑𝑑𝑑 = −𝑆𝑆𝑆𝑆𝑆𝑆 + 𝑉𝑉𝑉𝑉𝑉𝑉 + 𝜇𝜇𝜇𝜇𝜇𝜇 , we
have:
𝜕𝜕𝐺𝐺 𝜕𝜕𝐺𝐺
= −𝑆𝑆 and = 𝑉𝑉
𝜕𝜕𝑇𝑇 𝑃𝑃,𝑁𝑁 𝜕𝜕𝑃𝑃 𝑇𝑇,𝑁𝑁
• Thus, the volume and entropy are found by direct
differentiation of the Gibbs free energy 𝐺𝐺.
Thermodynamic Potentials
• D. Consider diffusive equilibrium between systems
involving different kinds of particles with different chemical
potentials.
• The Gibbs free energy of the total system is the sum over
all the different types of particles:
𝐺𝐺 = � 𝜇𝜇𝑖𝑖 𝑁𝑁𝑖𝑖
𝑖𝑖
• The condition that 𝐺𝐺 must be a minimum at equilibrium is
Δ𝐺𝐺 = � 𝜇𝜇𝑖𝑖 Δ𝑁𝑁𝑖𝑖 = 0
𝑖𝑖
at constant 𝑇𝑇 and 𝑃𝑃.
Thermodynamic Potentials
• Gibbs Free Energy and Phase Transition: For a
reversible isothermal, isobaric and non-diffusive
process we have 𝐺𝐺 = const.
• This is a particularly important result in
connection with processes involving change of
phase.
• For example, in a closed system, sublimation,
fusion, and vaporization take place isothermally
and isobarically.
• Hence, during such processes, the Gibbs free
energy of the system remains constant.
Thermodynamic Potentials
• Let us denote molar Gibbs free energy by
𝑔𝑔 = 𝐺𝐺/𝑛𝑛, where 𝑛𝑛 = the number of moles.
• Let, 𝑔𝑔′ = 𝑔𝑔𝑆𝑆 , 𝑔𝑔′′ = 𝑔𝑔𝐿𝐿 and 𝑔𝑔′′′ = 𝑔𝑔𝑉𝑉 denote the
molar Gibbs functions of a saturated solid,
saturated liquid, and saturated vapor,
respectively.
• Then the equation of the fusion curve is:
𝑔𝑔′ = 𝑔𝑔′′
• The equation for the vaporization curve is:
𝑔𝑔′′ = 𝑔𝑔′′′
Thermodynamic Potentials
• The equation for the sublimation curve is:
𝑔𝑔′ = 𝑔𝑔′′′
• At the triple point of a substance, any two of the
above equations hold
simultaneously to give:
𝑔𝑔′ = 𝑔𝑔′′ = 𝑔𝑔′′′
• The above determines
uniquely the constant
𝑇𝑇 and pressure 𝑃𝑃 at
the triple point.
Thermodynamic Potentials
• Gibbs Free Energy and Chemical Reactions:
The Gibbs function is extremely important in
chemistry, since chemical reactions begin and
end at the same equilibrium atmospheric
pressure and ambient temperature.
Thermodynamic Potentials
• Enthalpy- Physical Interpretation:
• Enthalpy is relevant and useful in the analysis of
isobaric processes that reach equilibrium (i.e. in
processes where 𝑑𝑑𝑑𝑑 = 0).
• A. Suppose we create a thermodynamic system
by bringing in the constituents in a thermally
insulated volume at constant pressure 𝑃𝑃.
• For example, we may create the system in a
chamber with movable piston that is exerting a
constant pressure 𝑃𝑃 while the walls and the
piston are heat insulators.
Thermodynamic Potentials
• We call the system to be in constant mechanical
contact with a ‘volume bath’ at constant pressure 𝑃𝑃.
• Thus while creating the system, besides supplying the
internal energy, we must also perform work ∫ 𝑃𝑃𝑃𝑃𝑃𝑃 =
𝑃𝑃 ∫ 𝑑𝑑𝑑𝑑 = 𝑃𝑃𝑃𝑃 in order to make room for the system to
reside.
• This extra work has to be done because the system is
being created against an applied constant pressure.
• Had we created the system in vacuum (with no applied
pressure on it), we would have needed only the
potential and kinetic energies of the particles, as parts
of the internal energy.
Thermodynamic Potentials
• The system would had no fixed volume (the
volume would have adjusted automatically).
• No extra work would have needed to create
the room for the system.
• Hence, the total amount of work done in
assembling such a system is:
𝑊𝑊� = 𝑈𝑈 + 𝑃𝑃𝑃𝑃
𝑃𝑃,𝑆𝑆
• The quantity 𝑈𝑈 + 𝑃𝑃𝑃𝑃 is called the enthalpy.
Thermodynamic Potentials
• B. Now, consider creation of the system from
scratch that is thermally interacting with the
environment while being created against a
constant pressure 𝑃𝑃.
• In this case the system will gain energy as heat
transferred from the environment (at constant
or variable 𝑇𝑇).
• The heat transferred will be part of the
internal energy gain.
Thermodynamic Potentials
• Again, we would need an extra work ∫ 𝑃𝑃𝑃𝑃𝑃𝑃 =
𝑃𝑃 ∫ 𝑑𝑑𝑑𝑑 = 𝑃𝑃𝑃𝑃 in order to make room for the
system to reside.
• Hence, the total amount of work done in
assembling such a system would be:
𝑊𝑊� = 𝑈𝑈 + 𝑃𝑃𝑃𝑃
𝑃𝑃
• Note that, the part 𝑃𝑃𝑃𝑃 is not a part of the
system’s internal energy.
• 𝑃𝑃𝑃𝑃 is the work done against external
environmental agents in order to make room.
Thermodynamic Potentials
• C. Consider two systems interacting thermally
and diffusively but each held at constant
pressure in thermally insulated chambers.
• Since, there is no external work done, and
since the system is thermally insulated, energy
or volume gained by one comes from the
other: 𝑑𝑑𝑈𝑈1 = −𝑑𝑑𝑈𝑈2 , 𝑑𝑑𝑉𝑉1 = −𝑑𝑑𝑉𝑉2
• Similarly, the particles gained by one are lost
by the other: 𝑑𝑑𝑁𝑁1 = −𝑑𝑑𝑁𝑁2
Thermodynamic Potentials
• The change of internal energy of one system will
have, in general two components:
Heat transferred: 𝑑𝑑𝑑𝑑
Work done: 𝑑𝑑𝑑𝑑
• Since, in this case the systems are allowed to
interact thermally, one will loose heat and the
other will gain, depending on their temperatures.
• One will do work on the other, depending on
their respective pressures.
Thermodynamic Potentials
• Thus the change in enthalpy in each system is
given by (𝑃𝑃1 , 𝑃𝑃2 individually constant):
𝑑𝑑𝐻𝐻𝑖𝑖 = 𝑑𝑑𝑈𝑈𝑖𝑖 + 𝑃𝑃𝑖𝑖 𝑑𝑑𝑉𝑉𝑖𝑖 + 𝑉𝑉𝑖𝑖 𝑑𝑑𝑃𝑃𝑖𝑖 = 𝑑𝑑𝑈𝑈𝑖𝑖 + 𝑃𝑃𝑖𝑖 𝑑𝑑𝑉𝑉𝑖𝑖
• The change of enthalpy of the combined system
is:
𝑑𝑑𝐻𝐻𝑡𝑡𝑡𝑡𝑡𝑡 = 𝑑𝑑𝑈𝑈1 + 𝑑𝑑𝑈𝑈2 + 𝑃𝑃1 𝑑𝑑𝑉𝑉1 + 𝑃𝑃2 𝑑𝑑𝑉𝑉2
⇒ 𝑑𝑑𝐻𝐻𝑡𝑡𝑡𝑡𝑡𝑡 = 0 + 𝑃𝑃1 − 𝑃𝑃2 𝑑𝑑𝑉𝑉1
• Since volume is gained by the system under
higher pressure, we get: if 𝑃𝑃1 > 𝑃𝑃2 ⇒ 𝑑𝑑𝑉𝑉1 > 0
and vice versa.
Thermodynamic Potentials
• Hence, 𝑑𝑑𝐻𝐻𝑡𝑡𝑡𝑡𝑡𝑡 > 0
• Hence, as the two systems approach
equilibrium under isobaric conditions,
changes in their enthalpy must be positive,
reaching a maximum at equilibrium.
The Porous Plug Experiment
• The Porous-Plug Experiment:
• Imagine a cylinder, thermally insulated and
equipped with two adiabatic pistons on opposite
sides of a porous wall that is also adiabatic.
• The importance of the porous wall is to permit
mass to flow from one chamber to another while
controlling the pressure, unlike a free expansion.
• The wall, shown ruled in horizontal lines, can be a
porous plug, a narrow constriction, or a series of
small holes.
The Porous Plug Experiment
• Between the left-hand piston and the wall there
is a gas at a pressure 𝑃𝑃𝑖𝑖 and a volume 𝑉𝑉𝑖𝑖 (𝑖𝑖 for
initial).
• Since the right-hand piston against the wall
prevents any gas from seeping through the
porous plug, the initial state of the gas is an
equilibrium state contained between the faces of
the two pistons.
• .
• .
• .

Seep= Flow or leak slowly through porous material or small holes.


The Porous Plug Experiment
• Now, imagine that both pistons move
simultaneously at different speeds to the right
such that a constant higher pressure 𝑃𝑃𝑖𝑖 is
maintained on the left-hand side of the porous
plug.
• A constant lower pressure 𝑃𝑃𝑓𝑓 is maintained on
the right-hand side.
• After all the gas has flowed through the porous
plug, the final equilibrium state of the system is
shown as below:
The Porous Plug Experiment
• There is no knowledge of the temperature of the
gas in either the initial state or the final state.
• The gas passes through dissipative non-
equilibrium states on its way from the initial
equilibrium state to the final equilibrium state.
• This process is known as a throttling process or a
Joule-Thomson expansion.

• .
The Porous Plug Experiment
• Although the intermediate states can not be
described by thermodynamic coordinates or
variables, we can draw an important conclusion
about the system.
• From the first law: Δ𝑈𝑈 = 𝑊𝑊𝑜𝑜𝑜𝑜 + 𝑄𝑄𝑖𝑖𝑖𝑖
where 𝑊𝑊𝑜𝑜𝑜𝑜 = net work done by the pistons
𝑄𝑄𝑖𝑖𝑖𝑖 = heat absorbed by the system.
• Since the chamber is thermally insulated, 𝑄𝑄𝑖𝑖𝑖𝑖 =
0.
• Hence, 𝑈𝑈𝑓𝑓 − 𝑈𝑈𝑖𝑖 = 𝑊𝑊𝑜𝑜𝑜𝑜
The Porous Plug Experiment
• The net work done by the pistons on the gas
causes the gas to flow across the boundary of the
system enclosing the porous plug:
0 𝑉𝑉𝑖𝑖
𝑊𝑊𝑜𝑜𝑜𝑜 = − � 𝑃𝑃𝑓𝑓 𝑑𝑑𝑑𝑑 − � 𝑃𝑃𝑖𝑖 𝑑𝑑𝑑𝑑 = −(𝑃𝑃𝑓𝑓 𝑉𝑉𝑓𝑓 − 𝑃𝑃𝑖𝑖 𝑉𝑉𝑖𝑖 )
𝑉𝑉𝑓𝑓 0
• The internal energies are different for the two
equilibrium end states of the Joule-Thomson
expansion.
• A state function can be devised for which there is
no difference in the end-states.
The Porous Plug Experiment
• Hence, we get:
𝑈𝑈𝑓𝑓 − 𝑈𝑈𝑖𝑖 = 𝑊𝑊𝑜𝑜𝑜𝑜 = −(𝑃𝑃𝑓𝑓 𝑉𝑉𝑓𝑓 − 𝑃𝑃𝑖𝑖 𝑉𝑉𝑖𝑖 )
⇒ 𝑈𝑈𝑖𝑖 + 𝑃𝑃𝑖𝑖 𝑉𝑉𝑖𝑖 = 𝑈𝑈𝑓𝑓 + 𝑃𝑃𝑓𝑓 𝑉𝑉𝑓𝑓
• Thus for the throttling process:
𝐻𝐻𝑖𝑖 = 𝐻𝐻𝑓𝑓
• In a throttling process the initial and final
enthalpies are equal.
• We can not say that throughout the process, the
enthalpy remains constant.
The Porous Plug Experiment
• Because the process involves non-equilibrium
intermediate stages in which enthalpy is not
defined.
• In plotting a throttling process on any
diagram, the initial and final equilibrium states
may be represented by points.
• The intermediate non-equilibrium states,
however, cannot be plotted.
Enthalpy
• Property-1: We have, 𝐻𝐻 = 𝑈𝑈 + 𝑃𝑃𝑃𝑃 ⇒ 𝑑𝑑𝑑𝑑 =
𝑑𝑑𝑑𝑑 + 𝑃𝑃𝑃𝑃𝑃𝑃 + 𝑉𝑉𝑉𝑉𝑉𝑉
• For a hydrostatic system, 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑 =
𝑑𝑑𝑑𝑑 − 𝑃𝑃𝑃𝑃𝑃𝑃 ⇒ 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑃𝑃𝑃𝑃𝑃𝑃
• Hence, 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑉𝑉𝑉𝑉𝑉𝑉
• Dividing both sides by 𝑑𝑑𝑑𝑑 we get:
𝑑𝑑𝑑𝑑 𝑑𝑑𝑑𝑑 𝑑𝑑𝑑𝑑
= + 𝑉𝑉
𝑑𝑑𝑑𝑑 𝑑𝑑𝑑𝑑 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑑𝑑𝑑𝑑
• At constant 𝑃𝑃: = + 0 = 𝐶𝐶𝑃𝑃
𝜕𝜕𝜕𝜕 𝑃𝑃 𝑑𝑑𝑑𝑑 𝑃𝑃
Enthalpy
• The state function enthalpy 𝐻𝐻 is related to an
experimental quantity, the isobaric heat capacity,
which is also a state function.
𝑓𝑓 𝑓𝑓
• Hence, 𝐻𝐻𝑓𝑓 − 𝐻𝐻𝑖𝑖 =
∫𝑖𝑖 𝑑𝑑𝑑𝑑 = ∫𝑖𝑖 𝐶𝐶𝑃𝑃 𝑑𝑑𝑑𝑑
• For example, for an ideal gas, 𝐶𝐶𝑃𝑃 is constant and
hence, 𝐻𝐻𝑓𝑓 − 𝐻𝐻𝑖𝑖 = 𝐶𝐶𝑃𝑃 (𝑇𝑇𝑓𝑓 − 𝑇𝑇𝑖𝑖 )
• Property-2: The change in enthalpy during an
isobaric process is equal to the heat that is
transferred between the system and the
𝑓𝑓
surroundings: 𝐻𝐻𝑓𝑓 − 𝐻𝐻𝑖𝑖 = ∫𝑖𝑖 𝑑𝑑𝑑𝑑 = 𝑄𝑄𝑃𝑃
Enthalpy
• For an isochoric (constant volume) process in a
hydrostatic system, heat is the flow of internal
energy.
• Whereas for an isobaric (constant pressure)
process in a hydrostatic system, heat is the flow
of enthalpy.
• The change of enthalpy of a system during an
isobaric chemical process is commonly called the
"heat of reaction," but the phrase enthalpy of
reaction is more informative.
Enthalpy
• Latent Enthalpy: If heat is added to the system
during a first-order phase transition (e.g.,
melting, boiling, or sublimation), then the change
of enthalpy of the system is called "latent heat."
• The word "latent" acknowledges that there is no
change in temperature of the system when
heating the system during a phase transition,
unlike heating without a phase transition.
• Again, it is more informative to use the phrase
latent enthalpy.
Enthalpy
• Property-4: Change of enthalpy during a reversible
adiabatic process:
𝑓𝑓 𝑓𝑓 𝑓𝑓
𝐻𝐻𝑓𝑓 − 𝐻𝐻𝑖𝑖 = � 𝑑𝑑𝑑𝑑 = � 𝑑𝑑𝑑𝑑 + 𝑉𝑉𝑉𝑉𝑉𝑉 = � 𝑉𝑉𝑉𝑉𝑉𝑉
𝑖𝑖 𝑖𝑖 𝑖𝑖
• This is equal to the area to the left of a curve for an
isentropic process on a 𝑃𝑃𝑃𝑃 diagram.
• Whereas the integral − ∫ 𝑃𝑃𝑃𝑃𝑃𝑃 is represented by the
area under an adiabatic curve on a 𝑃𝑃𝑃𝑃 diagram.
• There is a thermodynamic difference between the two
integrals.
Enthalpy
• The integral − ∫ 𝑃𝑃𝑃𝑃𝑃𝑃 is adiabatic work, which
changes the configuration of a system with
constant mass by changing the volume.
Enthalpy
• The integral ∫ 𝑉𝑉𝑉𝑉𝑉𝑉, known as the (negative)
flow-work in engineering practice, is energy
that is received by a flowing gas in a region of
higher pressure.
• For example, ∫ 𝑉𝑉𝑉𝑉𝑉𝑉 is the energy received
from a pump or piston, and then carried to a
region of lower pressure, such as in the
continuous Joule-Thomson expansion.
TdS Equations
• The entropy of a pure substance can be considered as a
function of any two variables, such as 𝑇𝑇 and 𝑉𝑉.
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝑉𝑉 𝑇𝑇
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
• Hence,𝑇𝑇𝑇𝑇𝑇𝑇 = 𝑇𝑇 𝑑𝑑𝑑𝑑 + 𝑇𝑇 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕 𝑑𝑑𝑑𝑑
• But 𝑇𝑇 = = 𝐶𝐶𝑉𝑉
𝜕𝜕𝜕𝜕 𝑉𝑉 𝑑𝑑𝑑𝑑 𝑉𝑉
• From Maxwell’s relation:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
=
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉
TdS Equations
• Hence we get the first 𝑇𝑇𝑇𝑇𝑇𝑇 – equation:
𝜕𝜕𝜕𝜕
𝑇𝑇𝑇𝑇𝑇𝑇 = 𝐶𝐶𝑉𝑉 𝑑𝑑𝑑𝑑 + 𝑇𝑇 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉
• Example-1: 1 mol of a van der Waals gas
undergoes a reversible isothermal expansion
from an initial molar volume 𝑣𝑣𝑖𝑖 to a final molar
volume 𝑣𝑣𝑓𝑓 . How much heat has been
transferred?
• Solution-1: For 1 mol, we use specific quantities.
TdS Equations
• Hence from the first 𝑇𝑇𝑇𝑇𝑇𝑇 equation:
𝜕𝜕𝜕𝜕
𝑇𝑇𝑇𝑇𝑇𝑇 = 𝑐𝑐𝑉𝑉 𝑑𝑑𝑑𝑑 + 𝑇𝑇 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉
• For a Van der Waal’s gas, we get
𝑅𝑅𝑅𝑅 𝑎𝑎
𝑃𝑃 = − 2
𝑣𝑣 − 𝑏𝑏 𝑣𝑣
𝜕𝜕𝜕𝜕 𝑅𝑅
⇒ =
𝜕𝜕𝜕𝜕 𝑉𝑉 𝑣𝑣 − 𝑏𝑏
𝑅𝑅 𝑅𝑅𝑅𝑅
• Hence, 𝑇𝑇𝑇𝑇𝑇𝑇 = 𝑐𝑐𝑉𝑉 𝑑𝑑𝑑𝑑 + 𝑇𝑇 𝑑𝑑𝑑𝑑 ⇒ 𝑇𝑇𝑇𝑇𝑇𝑇 = 𝑑𝑑𝑑𝑑 for
𝑣𝑣−𝑏𝑏 𝑣𝑣−𝑏𝑏
constant 𝑇𝑇
TdS Equations
• Since the process is reversible, 𝑞𝑞 = ∫ 𝑇𝑇𝑇𝑇𝑇𝑇
• Hence the heat transferred:
𝑣𝑣𝑓𝑓 𝑣𝑣𝑓𝑓 − 𝑏𝑏
𝑑𝑑𝑑𝑑
𝑞𝑞 = 𝑅𝑅𝑅𝑅 � = 𝑅𝑅𝑅𝑅 ln
𝑣𝑣𝑖𝑖 𝑣𝑣 − 𝑏𝑏 𝑣𝑣𝑖𝑖 − 𝑏𝑏
TdS Equations
• Second 𝑻𝑻𝑻𝑻𝑻𝑻 Equation: A second
𝑇𝑇𝑑𝑑𝑑𝑑 equation can be derived if the entropy of
a pure substance is regarded as a function
of 𝑇𝑇 and 𝑃𝑃.
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
• We have, 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝑃𝑃 𝑇𝑇
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
• Hence,𝑇𝑇𝑇𝑇𝑇𝑇 = 𝑇𝑇 𝑑𝑑𝑑𝑑 + 𝑇𝑇 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝑃𝑃 𝑇𝑇
𝜕𝜕𝜕𝜕 𝑑𝑑𝑑𝑑
• But, 𝑇𝑇 = = 𝐶𝐶𝑃𝑃
𝜕𝜕𝜕𝜕 𝑃𝑃 𝑑𝑑𝑑𝑑 𝑃𝑃
TdS Equations
• From Maxwell’s fourth relation:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
• =−
𝜕𝜕𝑃𝑃 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑃𝑃
• Hence,
𝜕𝜕𝑉𝑉
𝑇𝑇𝑇𝑇𝑇𝑇 = 𝐶𝐶𝑃𝑃 𝑑𝑑𝑇𝑇 − 𝑇𝑇 𝑑𝑑𝑑𝑑
𝜕𝜕𝑇𝑇 𝑃𝑃
which is called the second 𝑇𝑇𝑇𝑇𝑇𝑇 equation.
Gibbs-Helmholtz Equations
• The Gibbs-Helmholtz relations relate the temperature
derivatives of 𝐴𝐴 and 𝐺𝐺 to 𝑈𝑈 and 𝐻𝐻.
• Recall we have Maxwell’s relations (for 𝑁𝑁 = constant):
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = −𝑆𝑆𝑆𝑆𝑆𝑆 + 𝑉𝑉𝑉𝑉𝑉𝑉 ⇒ − =
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝜕𝜕𝑃𝑃
𝑑𝑑𝐴𝐴 = −𝑆𝑆𝑆𝑆𝑆𝑆 − 𝑃𝑃𝑃𝑃𝑃𝑃 ⇒ =
𝜕𝜕𝑉𝑉 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝑇𝑇 𝜕𝜕𝜕𝜕
𝑑𝑑𝐻𝐻 = 𝑇𝑇𝑑𝑑𝑆𝑆 + 𝑉𝑉𝑉𝑉𝑉𝑉 ⇒ =
𝜕𝜕𝜕𝜕 𝑆𝑆 𝜕𝜕𝑆𝑆 𝑃𝑃
𝜕𝜕𝑃𝑃 𝜕𝜕𝑇𝑇
𝑑𝑑𝑈𝑈 = 𝑇𝑇𝑑𝑑𝑆𝑆 − 𝑃𝑃𝑃𝑃𝑃𝑃 ⇒ − =
𝜕𝜕𝑆𝑆 𝑉𝑉 𝜕𝜕𝑃𝑃 𝑉𝑉
Gibbs-Helmholtz Equations
• Hence we have
𝜕𝜕 𝐺𝐺 1 1 𝐻𝐻
= −𝑆𝑆 − 𝐻𝐻 − 𝑇𝑇𝑇𝑇 2 = − 2
𝜕𝜕𝑇𝑇 𝑇𝑇 𝑇𝑇 𝑇𝑇 𝑇𝑇
𝑃𝑃
𝜕𝜕 𝐺𝐺 𝜕𝜕 𝐺𝐺
⇒ = −𝑇𝑇 2 = 𝐻𝐻
𝜕𝜕 1/𝑇𝑇 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝑃𝑃
𝑃𝑃
𝜕𝜕 𝐴𝐴 1 𝜕𝜕𝜕𝜕 𝐴𝐴 𝑆𝑆 𝑈𝑈−𝑇𝑇𝑇𝑇 𝑈𝑈
= − = − − = − 2
𝜕𝜕𝜕𝜕 𝑇𝑇 𝑉𝑉 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉 𝑇𝑇 2 𝑇𝑇 𝑇𝑇 2 𝑇𝑇
𝜕𝜕 𝐴𝐴 𝜕𝜕 𝐴𝐴
⇒ = −𝑇𝑇 2 = 𝑈𝑈
𝜕𝜕 1/𝑇𝑇 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇 𝑉𝑉
𝑉𝑉
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
TdS Equations
• For a pure substance, the entropy may be
considered as a function of any two of the
variables (𝑃𝑃, 𝑉𝑉, 𝑇𝑇):
𝑆𝑆 = 𝑆𝑆(𝑇𝑇, 𝑉𝑉) or 𝑆𝑆 = 𝑆𝑆 𝑇𝑇, 𝑃𝑃 or 𝑆𝑆 = 𝑆𝑆(𝑃𝑃, 𝑉𝑉)
• These dependence of the entropy together
with the Maxwell’s relations gives us
expressions for the heat transferred in the
form of 𝑑𝑑𝑑𝑑 = 𝑇𝑇𝑇𝑇𝑇𝑇 under certain constraints.
• These relations ae called the 𝑇𝑇𝑇𝑇𝑇𝑇 equations.
TdS Equations
• The First 𝑻𝑻𝑻𝑻𝑻𝑻 Equation: Let the entropy be a function
of 𝑇𝑇 and 𝑉𝑉, i.e. 𝑆𝑆 = 𝑆𝑆(𝑇𝑇, 𝑉𝑉) :
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝑉𝑉 𝑇𝑇
• But, from 𝑑𝑑𝑑𝑑 = 𝑇𝑇𝑇𝑇𝑇𝑇 at constant 𝑉𝑉, we get
𝑑𝑑𝑑𝑑 𝑇𝑇𝑇𝑇𝑇𝑇|𝑉𝑉 𝜕𝜕𝜕𝜕
𝐶𝐶𝑉𝑉 = = = 𝑇𝑇
𝑑𝑑𝑑𝑑 𝑉𝑉 𝑑𝑑𝑑𝑑|𝑉𝑉 𝜕𝜕𝜕𝜕 𝑉𝑉
which is the heat capacity at constant volume.
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
• Again, from Maxwell’s relation: =
𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉
TdS Equations
• Hence we have,
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑇𝑇𝑇𝑇𝑇𝑇 = 𝑇𝑇 𝑑𝑑𝑑𝑑 + 𝑇𝑇 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑇𝑇
𝜕𝜕𝜕𝜕
= 𝐶𝐶𝑉𝑉 𝑑𝑑𝑑𝑑 + 𝑇𝑇 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉
• This is called the first 𝑇𝑇𝑇𝑇𝑇𝑇 equation.
• This is useful in evaluating the heat transferred in
different processes, e.g reversible isothermal or
isochoric processes.
TdS Equations
• The Second 𝑻𝑻𝑻𝑻𝑻𝑻 Equation: Let the entropy be a
function of 𝑇𝑇 and 𝑃𝑃, i.e. 𝑆𝑆 = 𝑆𝑆(𝑇𝑇, 𝑃𝑃) :
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑃𝑃
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝑃𝑃 𝑇𝑇
• But, from 𝑑𝑑𝑑𝑑 = 𝑇𝑇𝑇𝑇𝑇𝑇 at constant 𝑃𝑃, we get
𝑑𝑑𝑑𝑑 𝑇𝑇𝑇𝑇𝑇𝑇|𝑃𝑃 𝜕𝜕𝜕𝜕
𝐶𝐶𝑃𝑃 = = = 𝑇𝑇
𝑑𝑑𝑑𝑑 𝑃𝑃 𝑑𝑑𝑑𝑑|𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃
which is the heat capacity at constant volume.
𝜕𝜕𝜕𝜕 𝜕𝜕𝑉𝑉
• Again, from Maxwell’s relation: =−
𝜕𝜕𝑃𝑃 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑃𝑃
TdS Equations
• Hence we have,
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑇𝑇𝑇𝑇𝑇𝑇 = 𝑇𝑇 𝑑𝑑𝑑𝑑 + 𝑇𝑇 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝑃𝑃 𝑇𝑇
𝜕𝜕𝑉𝑉
= 𝐶𝐶𝑃𝑃 𝑑𝑑𝑑𝑑 − 𝑇𝑇 𝑑𝑑𝑃𝑃
𝜕𝜕𝜕𝜕 𝑃𝑃
• This is called the second 𝑇𝑇𝑇𝑇𝑇𝑇 equation.
• This is useful in evaluating the heat transferred in
different processes, e.g reversible isothermal or
isobaric processes.
TdS Equations
• The second 𝑇𝑇𝑇𝑇𝑇𝑇 equation is more useful than
the first 𝑇𝑇𝑇𝑇𝑇𝑇 equation because the partial
derivative is evaluated at constant pressure
rather than at constant volume (which does
not happen often).
• Applications of the Second 𝑇𝑇𝑇𝑇𝑇𝑇 Equation:
• A. Reversible Isothermal Change of Pressure:
𝜕𝜕𝜕𝜕
At constant 𝑇𝑇: 𝑇𝑇𝑇𝑇𝑇𝑇 = −𝑇𝑇 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑃𝑃
TdS Equations
• Integrating, the heat transferred becomes:
𝜕𝜕𝜕𝜕
𝑄𝑄 = � 𝑇𝑇𝑇𝑇𝑇𝑇 = −𝑇𝑇 � 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑃𝑃
• Recalling the definition of volume expansivity:
1 𝜕𝜕𝜕𝜕
𝛽𝛽 =
𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃
we get, 𝑄𝑄 = −𝑇𝑇 ∫ 𝑉𝑉𝑉𝑉 𝑑𝑑𝑑𝑑
• The above can be integrated when the dependence of
𝑉𝑉 and 𝛽𝛽 on the pressure 𝑃𝑃 is known.
TdS Equations
• In the case of a solid or liquid, neither 𝑉𝑉 nor 𝛽𝛽
is very sensitive to a change of pressure.
• For example, in mercury (Hg), as the pressure
is increased from zero to a thousand times
the atmospheric pressure at room
temperature:
• A. Volume 𝑉𝑉 increases by only 0.333% and
• B. Volume expansivity 𝛽𝛽 changes by only 4%.
TdS Equations
• Hence, the heat transferred during reversible
isothermal change of pressure for substances having 𝑉𝑉
and 𝛽𝛽 independent of pressure:
𝑃𝑃𝑓𝑓
𝑄𝑄 =−𝑇𝑇𝑇𝑇𝑇𝑇 ∫𝑃𝑃 𝑑𝑑𝑑𝑑 = −𝑇𝑇𝑇𝑇𝑇𝑇(𝑃𝑃𝑓𝑓 − 𝑃𝑃𝑖𝑖 )
𝑖𝑖
• If 𝛽𝛽 > 0, as the pressure is increased isothermally, heat
will flow OUT.
• If 𝛽𝛽 < 0, as the pressure is increased isothermally, heat
will be ABSORBED.
• This is the case for water between 0∘ C and 4∘ C, where
the volume expansivity is negative. As the pressure is
increased isothermally, heat is absorbed by water.
TdS Equations
• B. Reversible Adiabatic Change of Pressure: At
𝜕𝜕𝜕𝜕
constant 𝑆𝑆: 𝑇𝑇𝑇𝑇𝑇𝑇 = 0 = 𝐶𝐶𝑃𝑃 𝑑𝑑𝑑𝑑 − 𝑇𝑇 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑃𝑃
𝑇𝑇 𝜕𝜕𝜕𝜕 𝑇𝑇𝑇𝑇𝑇𝑇
⇒ 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑
𝐶𝐶𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃 𝐶𝐶𝑃𝑃
• In the case of a solid or liquid, an increase of
pressure of as much as 1000 atm produces only a
small temperature change.
• Also, experiment shows that 𝐶𝐶𝑝𝑝 hardly changes,
even for an increase of 10,000 atm.
TdS Equations
• Hence, for a solid of liquid,
𝑇𝑇𝑇𝑇𝑇𝑇
Δ𝑇𝑇 ≈ (𝑃𝑃𝑓𝑓 − 𝑃𝑃𝑖𝑖 )
𝐶𝐶𝑃𝑃
• The Third 𝑻𝑻𝑻𝑻𝑻𝑻 Equation: Let the entropy be a function
of 𝑃𝑃 and 𝑉𝑉, i.e. 𝑆𝑆 = 𝑆𝑆(𝑃𝑃, 𝑉𝑉) :
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃
• Applying chain rule, we get:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃
TdS Equations
• From 𝑑𝑑𝑑𝑑 = 𝑇𝑇𝑇𝑇𝑇𝑇 at constant 𝑉𝑉, we get
𝑑𝑑𝑑𝑑 𝑇𝑇𝑇𝑇𝑇𝑇|𝑉𝑉 𝜕𝜕𝜕𝜕
𝐶𝐶𝑉𝑉 = = = 𝑇𝑇
𝑑𝑑𝑑𝑑 𝑉𝑉 𝑑𝑑𝑑𝑑|𝑉𝑉 𝜕𝜕𝜕𝜕 𝑉𝑉
which is the heat capacity at constant volume.
• From 𝑑𝑑𝑑𝑑 = 𝑇𝑇𝑇𝑇𝑇𝑇 at constant 𝑃𝑃, we get
𝑑𝑑𝑑𝑑 𝑇𝑇𝑇𝑇𝑇𝑇|𝑃𝑃 𝜕𝜕𝜕𝜕
𝐶𝐶𝑃𝑃 = = = 𝑇𝑇
𝑑𝑑𝑑𝑑 𝑃𝑃 𝑑𝑑𝑑𝑑|𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃
which is the heat capacity at constant pressure.
TdS Equations
• Hence we get,
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑇𝑇𝑇𝑇𝑇𝑇 = 𝐶𝐶𝑉𝑉 𝑑𝑑𝑑𝑑 + 𝐶𝐶𝑃𝑃 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑃𝑃
which is the third 𝑇𝑇𝑇𝑇𝑇𝑇 equation.
• Equivalent forms of 𝑻𝑻𝑻𝑻𝑻𝑻 Equations: Recall the
definitions of volume expansivity 𝛽𝛽 and the
isothermal compressibility 𝜅𝜅:
1 𝜕𝜕𝜕𝜕 1 𝜕𝜕𝜕𝜕
𝛽𝛽 = , 𝜅𝜅 = −
𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑇𝑇
TdS Equations
• From the first 𝑇𝑇𝑇𝑇𝑇𝑇 equation:
𝜕𝜕𝜕𝜕
𝑇𝑇𝑇𝑇𝑇𝑇 = 𝐶𝐶𝑉𝑉 𝑑𝑑𝑑𝑑 + 𝑇𝑇 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉
• But, from chain rule, and the rule for inverse:
𝛽𝛽 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝑃𝑃 𝜕𝜕𝜕𝜕
=− / =− =
𝜅𝜅 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝑉𝑉 𝑇𝑇 𝜕𝜕𝜕𝜕 𝑉𝑉
• Hence, the first 𝑇𝑇𝑇𝑇𝑇𝑇 equation becomes:
𝑇𝑇𝑇𝑇
𝑇𝑇𝑇𝑇𝑇𝑇 = 𝐶𝐶𝑉𝑉 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝜅𝜅
TdS Equations
• The second 𝑇𝑇𝑇𝑇𝑇𝑇 equation can be written as:
𝜕𝜕𝜕𝜕
𝑇𝑇𝑇𝑇𝑇𝑇 = 𝐶𝐶𝑃𝑃 𝑑𝑑𝑑𝑑 − 𝑇𝑇 𝑑𝑑𝑑𝑑 = 𝐶𝐶𝑃𝑃 𝑑𝑑𝑑𝑑 − 𝑇𝑇𝑇𝑇𝑇𝑇𝑇𝑇𝑇𝑇
𝜕𝜕𝜕𝜕 𝑃𝑃
• The third 𝑇𝑇𝑇𝑇𝑇𝑇 equation is:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑇𝑇𝑇𝑇𝑇𝑇 = 𝐶𝐶𝑉𝑉 𝑑𝑑𝑑𝑑 + 𝐶𝐶𝑃𝑃 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
• Using, 𝛽𝛽/𝜅𝜅 = ,
𝑉𝑉𝑉𝑉 = we get:
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃
𝐶𝐶𝑉𝑉 𝜅𝜅 𝐶𝐶𝑃𝑃
𝑇𝑇𝑇𝑇𝑇𝑇 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑑𝑑
𝛽𝛽 𝑉𝑉𝑉𝑉
Heat Capacity Equations
• From the first and second 𝑇𝑇𝑇𝑇𝑇𝑇 equations:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑇𝑇𝑇𝑇𝑇𝑇 = 𝐶𝐶𝑉𝑉 𝑑𝑑𝑑𝑑 + 𝑇𝑇 𝑑𝑑𝑑𝑑 , 𝑇𝑇𝑇𝑇𝑇𝑇 = 𝐶𝐶𝑃𝑃 𝑑𝑑𝑑𝑑 − 𝑇𝑇 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃
• To solve for 𝑑𝑑𝑑𝑑, we get:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉 𝑑𝑑𝑑𝑑 = 𝑇𝑇 𝑑𝑑𝑑𝑑 + 𝑇𝑇 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑃𝑃
• But from 𝑇𝑇 = 𝑇𝑇(𝑉𝑉, 𝑃𝑃):
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝑑𝑑𝑑𝑑 = 𝑑𝑑𝑑𝑑 + 𝑑𝑑𝑃𝑃
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝑃𝑃 𝑉𝑉
Heat Capacity Equations
• Comparing the coefficients, we get:
𝜕𝜕𝜕𝜕
𝑇𝑇
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑉𝑉
=
𝜕𝜕𝜕𝜕 𝑃𝑃 𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉
𝜕𝜕𝜕𝜕
𝑇𝑇
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑃𝑃
=
𝜕𝜕𝜕𝜕 𝑉𝑉 𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉
• From both the above equations, we get:
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉 = 𝑇𝑇
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃
Heat Capacity Equations
• Using the result of partial derivatives of three variables:
𝜕𝜕𝜕𝜕 𝜕𝜕𝑇𝑇 𝜕𝜕𝜕𝜕
= −1
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝑉𝑉 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇
we get,
𝜕𝜕𝜕𝜕 𝜕𝜕𝑉𝑉 𝜕𝜕𝑃𝑃
=−
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝑇𝑇 𝑃𝑃 𝜕𝜕𝑉𝑉 𝑇𝑇
• Hence, the heat capacity equation becomes:
2
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉 = 𝑇𝑇 = −𝑇𝑇
𝜕𝜕𝜕𝜕 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝑉𝑉 𝑇𝑇
Heat Capacity Equations
• Conclusions form the Heat Capacity Equation: The
heat capacity equation gives important physical
consequences:
• A. Since 𝜕𝜕𝜕𝜕/𝜕𝜕𝜕𝜕 𝑇𝑇 > 0 i.e. always positive, for all
known substances, and since 𝜕𝜕𝜕𝜕/𝜕𝜕𝜕𝜕 2𝑃𝑃 > 0, we get:
𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉 ≥ 0
or 𝐶𝐶𝑃𝑃 can never be less than 𝐶𝐶𝑉𝑉 .
• B. As 𝑇𝑇 → 0, 𝐶𝐶𝑃𝑃 → 𝐶𝐶𝑉𝑉 or, at absolute zero, the two
heat capacities are equal.
• C. 𝐶𝐶𝑃𝑃 = 𝐶𝐶𝑉𝑉 , when 𝜕𝜕𝜕𝜕/𝜕𝜕𝜕𝜕 𝑃𝑃 = 0. For example, at 4∘ C,
the temperature at which the density of water is a
maximum, 𝐶𝐶𝑃𝑃 = 𝐶𝐶𝑉𝑉 .
Heat Capacity Equations
• Laboratory measurements of the heat capacity of
solids and liquids usually take place at constant
pressure and for unit mass.
• Also, it also very difficult to measure directly the
specific heat at constant volume for solid of
liquids.
• Therefore, the data is reported in terms of
specific heat 𝑐𝑐𝑝𝑝 .
• The equation for the difference in the specific
heats is very useful in calculating 𝑐𝑐𝑉𝑉 in terms of
𝑐𝑐𝑃𝑃 and other measurable quantities.
Heat Capacity Equations
• Recalling the definitions of volume expansivity 𝛽𝛽 and
the isothermal compressibility 𝜅𝜅:
1 𝜕𝜕𝜕𝜕 1 𝜕𝜕𝜕𝜕
𝛽𝛽 = , 𝜅𝜅 = −
𝑉𝑉 𝜕𝜕𝜕𝜕 𝑃𝑃 𝑉𝑉 𝜕𝜕𝜕𝜕 𝑇𝑇
we get,
2
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕
𝐶𝐶𝑃𝑃 − 𝐶𝐶𝑉𝑉 = 𝑇𝑇 = −𝑇𝑇
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑇𝑇
2
𝜕𝜕𝜕𝜕 𝜕𝜕𝑣𝑣 𝜕𝜕𝑣𝑣 𝜕𝜕𝜕𝜕
⇒ 𝑐𝑐𝑃𝑃 − 𝑐𝑐𝑉𝑉 = 𝑇𝑇 = −𝑇𝑇
𝜕𝜕𝜕𝜕 𝑉𝑉
𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝜕𝜕 𝑃𝑃
𝜕𝜕𝑣𝑣 𝑇𝑇
Heat Capacity Equations
• Hence,
2
1 𝜕𝜕𝑉𝑉
2
𝜕𝜕𝜕𝜕 𝜕𝜕𝜕𝜕 𝑉𝑉 2 𝜕𝜕𝜕𝜕 𝑃𝑃
𝑐𝑐𝑃𝑃 − 𝑐𝑐𝑉𝑉 = −𝑇𝑇 = 𝑇𝑇𝑇𝑇
𝜕𝜕𝜕𝜕 𝑃𝑃 𝜕𝜕𝜕𝜕 𝑇𝑇 1 𝜕𝜕𝑉𝑉

𝑉𝑉 𝜕𝜕𝑃𝑃 𝑇𝑇
𝛽𝛽 2
𝑐𝑐𝑃𝑃 − 𝑐𝑐𝑉𝑉 = 𝑇𝑇𝑇𝑇
𝜅𝜅
Thermal Physics

Instructor-
Golam Dastegir Al-Quaderi
Associate Professor
Department of Physics, DU
Phase Transition
• A typical phase diagram of a 𝑃𝑃 − 𝑉𝑉 − 𝑇𝑇 system is
shown as below:
• .
• .

• The solid lines are boundaries between phases. These


lines are called coexistence curves.
Phase Transition
• 𝑷𝑷𝑷𝑷𝑷𝑷 Surface: The equation of state for a single
component system may be written as:
𝑓𝑓 𝑃𝑃, 𝑉𝑉, 𝑇𝑇 = 0
• The single constraint 𝑓𝑓 𝑃𝑃, 𝑉𝑉, 𝑇𝑇 = 0 on the three
state variables defines a surface in 𝑃𝑃𝑃𝑃𝑃𝑃- space.
• This is called the 𝑃𝑃𝑃𝑃𝑃𝑃-surface.
• The equation of the constraint may in principle be
inverted to yield 𝑃𝑃 = 𝑃𝑃(𝑉𝑉, 𝑇𝑇) or 𝑉𝑉 = 𝑉𝑉(𝑇𝑇, 𝑃𝑃) or
𝑇𝑇 = 𝑇𝑇(𝑃𝑃, 𝑉𝑉).
Phase Transition
• 𝑃𝑃𝑃𝑃𝑃𝑃-surface for an ideal gas.
Phase Transition
• Real PVT- surfaces are much richer than that
for the ideal gas, because real systems
undergo phase transitions.
Phase Transition
• We can take projections of the PVT surfaces
on the 𝑉𝑉𝑉𝑉-, 𝑉𝑉𝑉𝑉- and 𝑇𝑇𝑇𝑇-planes:
Phase Transition
• For example, we have the projection of the
PVT-surface of a real gas on the PV-plane as
below:
Phase Transition
• The high temperature isotherms resemble those
of the ideal gas.
• But as one cools below the critical temperature
𝑇𝑇𝐶𝐶 , the isotherms become singular.
• Precisely at 𝑇𝑇 = 𝑇𝑇𝐶𝐶 , the isotherm 𝑃𝑃 = 𝑃𝑃(𝑣𝑣, 𝑇𝑇𝐶𝐶 )
becomes perfectly horizontal at 𝑣𝑣 = 𝑣𝑣𝐶𝐶 , which is
the critical molar volume.
• Thus the isothermal compressibility 𝜅𝜅 𝑇𝑇 =
− 1/𝑉𝑉 𝜕𝜕𝜕𝜕/𝜕𝜕𝜕𝜕 𝑇𝑇 diverges at 𝑇𝑇 = 𝑇𝑇𝐶𝐶 .
Phase Transition
• Below 𝑇𝑇𝐶𝐶 , the isotherms have a flat portion,
corresponding to a two-phase region where
liquid and vapor coexist.
• For water we have
on the PT-plane the
projection of the
PVT-surface is:
Phase Transition
• liquid-vapor phase coexistence occurs along a curve, called
the vaporization (or boiling) curve.
• The density changes discontinuously across this curve; for
H2O, the liquid is approximately 1000 times denser than
the vapor at atmospheric pressure.
• The density discontinuity vanishes at the critical point.
• Note that, one can continuously transform between liquid
and vapor phases, without encountering any phase
transitions, by going around the critical point and avoiding
the two-phase region.
• During Phase transitions, thermodynamic properties are
singular or discontinuous along certain curves on the PVT-
surface.
Phase Transition
• Order of Phase Transition: The order of the
singularity of the thermodynamic potentials is
often taken as a classification of the phase
transition.
• If the thermodynamic potentials 𝐸𝐸, 𝐹𝐹, 𝐺𝐺, and
𝐻𝐻 have discontinuous or divergent m-th
derivatives, the transition between the
respective phases is said to be m-th order .
Phase Transition
• We may characterize the familiar phase transitions by
either of the following equivalent statements:
• (A) There are changes of molar entropy and of molar
volume.
• (B) The first-order derivatives of the molar Gibbs
function change discontinuously.
• First order Phase Transition: Any phase change that
satisfies these requirements is known as a phase
change of the first order.
• For such a phase change, the temperature variations of
𝑔𝑔, 𝑠𝑠, 𝑣𝑣, and 𝐶𝐶𝐶𝐶 are shown in the figure:
Phase Transition
• First order phase transition;
Clausius-Clapeyron Equation
• The second TdS equation provides an
indeterminate result when applied to a first-
order phase transition.
• For 1 mol, we have
𝑇𝑇𝑇𝑇𝑇𝑇 = 𝑐𝑐𝑃𝑃 𝑑𝑑𝑑𝑑 − 𝑇𝑇𝑇𝑇𝑇𝑇𝑇𝑇𝑇𝑇
where 𝑐𝑐𝑃𝑃 = ∞ and 𝑑𝑑𝑑𝑑 = 0, while 𝛽𝛽 = ∞ and
𝑑𝑑𝑑𝑑 = 0.
• The first T dS equation, however, may be
integrated through the phase transition.
Clausius-Clapeyron Equation
• When 1 mol of substance is converted reversibly,
isothermally, and isobarically from phase (i) to phase
(f), the first TdS equation is
𝜕𝜕𝜕𝜕
𝑇𝑇𝑇𝑇𝑇𝑇 = 𝑐𝑐𝑉𝑉 𝑑𝑑𝑑𝑑 + 𝑇𝑇 𝑑𝑑𝑑𝑑
𝜕𝜕𝜕𝜕 𝑉𝑉
• This may be integrated with the understanding that the
various P's and T's at which a phase transition occurs
obey a relation in which P is a function of T only,
independent of V.
𝜕𝜕𝜕𝜕 𝑑𝑑𝑑𝑑
• Hence, =
𝜕𝜕𝜕𝜕 𝑉𝑉 𝑑𝑑𝑑𝑑
Clausius-Clapeyron Equation
• Hence we have,
𝑓𝑓
𝑓𝑓 𝑖𝑖
𝑑𝑑𝑑𝑑 𝑓𝑓
� 𝑇𝑇𝑇𝑇𝑇𝑇 = 𝑇𝑇 𝑠𝑠 − 𝑠𝑠 = 𝑇𝑇 (𝑣𝑣 − 𝑣𝑣 𝑖𝑖 )
𝑖𝑖 𝑑𝑑𝑑𝑑
𝑑𝑑𝑑𝑑 𝑠𝑠 𝑓𝑓 − 𝑠𝑠 𝑖𝑖 1 𝑇𝑇(𝑠𝑠 𝑓𝑓 − 𝑠𝑠 𝑖𝑖 )
= 𝑓𝑓 𝑖𝑖
=
𝑑𝑑𝑑𝑑 𝑣𝑣 − 𝑣𝑣 𝑇𝑇 𝑣𝑣 𝑓𝑓 − 𝑣𝑣 𝑖𝑖
• But 𝑑𝑑𝑑𝑑 = 𝑉𝑉𝑉𝑉𝑉𝑉 + 𝑇𝑇𝑇𝑇𝑇𝑇 which for molar
amount: 𝑑𝑑𝑑 = 𝑣𝑣𝑣𝑣𝑣𝑣 + 𝑇𝑇𝑇𝑇𝑇𝑇
• At isobaric case, 𝑑𝑑𝑑 = 𝑇𝑇𝑇𝑇𝑇𝑇
Clausius-Clapeyron Equation
• Hence we have Clasius-Clapeyron equation:
𝑑𝑑𝑑𝑑 1 ℎ 𝑓𝑓 − ℎ𝑖𝑖
==
𝑑𝑑𝑑𝑑 𝑇𝑇 𝑣𝑣 𝑓𝑓 − 𝑣𝑣 𝑖𝑖
• This applies to any first order change of phase
or any transition that occurs at constant T and
P.
• The difference in molar enthalpies at a fixed
pressure ℎ 𝑓𝑓 − ℎ𝑖𝑖 is also known as the molar
latent heat.
Clausius-Clapeyron Equation
• Alternative Derivation of the Clasius-
Clapeyron Equation:
• We have, the Gibbs function remains constant
during a reversible process taking place at
constant temperature and pressure.
• Hence, for a change of phase at 𝑇𝑇 and 𝑃𝑃, the
molar (or specific) Gibbs functions are equal
at the two end-states:
𝑔𝑔𝑖𝑖 = 𝑔𝑔 𝑓𝑓
Clausius-Clapeyron Equation
• For a phase change at 𝑇𝑇 + 𝑑𝑑𝑑𝑑 and 𝑃𝑃 + 𝑑𝑑𝑑𝑑:
𝑔𝑔𝑖𝑖 + 𝑑𝑑𝑔𝑔𝑖𝑖 = 𝑔𝑔 𝑓𝑓 + 𝑑𝑑𝑔𝑔 𝑓𝑓
• Subtracting we get,
𝑑𝑑𝑔𝑔𝑖𝑖 = 𝑑𝑑𝑔𝑔 𝑓𝑓
⇒ −𝑠𝑠 𝑖𝑖 𝑑𝑑𝑑𝑑 + 𝑣𝑣 𝑖𝑖 𝑑𝑑𝑑𝑑 = −𝑠𝑠 𝑓𝑓 𝑑𝑑𝑑𝑑 + 𝑣𝑣 𝑓𝑓 𝑑𝑑𝑑𝑑
𝑑𝑑𝑑𝑑 𝑠𝑠 𝑓𝑓 − 𝑠𝑠 𝑖𝑖
⇒ = 𝑓𝑓
𝑑𝑑𝑑𝑑 𝑣𝑣 − 𝑣𝑣 𝑖𝑖
• Hence we get:
𝑑𝑑𝑑𝑑 𝑠𝑠 𝑓𝑓 − 𝑠𝑠 𝑖𝑖 1 𝑇𝑇(𝑠𝑠 𝑓𝑓 − 𝑠𝑠 𝑖𝑖 )
= 𝑓𝑓 𝑖𝑖
=
𝑑𝑑𝑑𝑑 𝑣𝑣 − 𝑣𝑣 𝑇𝑇 𝑣𝑣 𝑓𝑓 − 𝑣𝑣 𝑖𝑖

You might also like