1.2 The Zeroth Law of Thermodynamics: Thermodynamic Laws
1.2 The Zeroth Law of Thermodynamics: Thermodynamic Laws
(1.2.2b)f1(P1,V1)=f3(P3,V3).D
strictly single valued; it exhibits non-uniqueness because of both the way it is defined and the
properties of real thermometers. For example, two laboratories’ temperature measurements may
differ by as much as 2 mK around 400 K, yet both comply with the ITS-90, assuming other
uncertainties are negligible. Therefore, at this level of accuracy, measured thermodynamic properties
may not appear to be smooth functions.
This chapter introduces high precision thermometry for those requiring a close match to the
thermodynamic temperature. To achieve the highest accuracies close adherence to the published
guidelines [2] is necessary. Lower accuracy thermometry is covered in other publications and
guidelines [4–7]. Since it is not possible to cover all thermometry applications for all possible
environments; in this chapter the emphasis is on making measurements traceable to the ITS-90. In
particular, the limits on accuracy and precision are examined in detail. Unless otherwise stated, all
uncertainties are reported as the standard uncertainty or one standard deviation.
At the extremes of temperature, the use or ITS-90 may not always be appropriate because new
techniques for realising the temperature scale are constantly developed. Extensions of thermometry
to very high and very low temperatures are outlined.
CP∆T = CV∆T + R ∆T
CP = CV + R
C P – CV = R
Difference Between Isothermal and Adiabatic Process
An isothermal process is a process that occurs under constant temperature but other parameters of
the system can be changed accordingly. On the other hand, in an adiabatic process, heat transfer
occurs to keep the temperature constant. The main difference between isothermal and adiabatic
process is that the isothermal process occurs under constant temperature, while the adiabatic
process occurs under varying temperature. The work done in an isothermal process is due to the
change in the net heat content of the system. Meanwhile, the work done in an adiabatic process is
due to the change in its internal energy.
Clausius’s Statement
It is impossible to construct a device operating in a cycle that can transfer heat from a colder body to
a warmer one without consuming any work. Also, energy will not flow spontaneously from a low-
temperature object to a higher-temperature object. It is important to note that we are referring to
the net transfer of energy. Energy transfer can take place from a cold object to a hot object by the
transfer of energetic particles or electromagnetic radiation. However, the net transfer will occur from
the hot object to the cold object in any spontaneous process. And some form of work is needed to
transfer the net energy to the hot object. In other words, unless the compressor is driven by an
external source, the refrigerator won’t be able to operate. The heat pump and refrigerator work on
Clausius’s statement.
Both Clausius’s and Kelvin’s statements are equivalent, i.e., a device violating Clausius’s statement
will also violate Kelvin’s statement and vice versa
7
Each of the four distinct processes are reversible. Using the fact that no heat enters or leaves in
adiabatic processes we can show that the work done in one cycle, W = Q 1 - Q3 where Q1 is the heat
entering at tempertature TH in the isothermal process A -> B and Q3 is the heat leaving at
temperature TC in the isothermal process C -> D.
By using the ideal gas equation (pV = nRT), the fact that W = Q for isothermal processes
and the fact that for adiabatic ideal gas processes it can be shown
that,
Remember, this is the ideal heat engine (reversible) efficiency. It sets the maximum theoretically
attainable efficiency of any real engine operating between the same two temperatures.
8
Be careful. The temperatures in the ideal gas law must be in Kelvin, therefore the temperatures in
the efficiency equation are also in Kelvin.
Entropy Changes in Reversible Processes
A change is said to occur reversibly when it can be carried out in a series of infinitesimal steps, each
one of which can be undone by making a similarly minute change to the conditions that bring the
change about. For example, the reversible expansion of a gas can be achieved by reducing the
external pressure in a series of infinitesimal steps; reversing any step will restore the system and the
surroundings
to their previous state. Similarly, heat can be transferred reversibly between two bodies by changing
the temperature difference between them in infinitesimal steps each of which can be undone by
reversing the temperature difference.
The most widely cited example of an irreversible change is the free expansion of a gas into a vacuum.
Although the system can always be restored to its original state by recompressing the gas, this would
require that the
surroundings
perform work on the gas. Since the gas does no work on the surrounding in a free expansion (the
external pressure is zero, so PΔV=0PΔV=0,) there will be a permanent change in the
surroundings
. Another example of irreversible change is the conversion of
mechanical work
into frictional heat; there is no way, by reversing the motion of a
weight
along a surface, that the heat released due to friction can be restored to the system.
hese diagrams show the same expansion and compression ±ΔV carried out in different numbers of
steps ranging from a single step at the top to an "infinite" number of steps at the bottom. As the
number of steps increases, the processes become less irreversible; that is, the difference between
the work done in expansion and that required to re-compress the gas diminishes. In the limit of an
”infinite” number of steps (bottom), these work terms are identical, and both the system and
surroundings
(the “world”) are unchanged by the expansion-compression cycle. In all other cases the system (the
gas) is restored to its initial state, but the
surroundings
are forever changed.
Work and Reversibility
Changes in entropy (ΔSΔS), together with changes in enthalpy (ΔHΔH), enable us to predict in which
direction a chemical or
physical change
will occur spontaneously. Before discussing how to do so, however, we must understand the
difference between a reversible process and an irreversible one. In a reversible process, every
intermediate
state between the extremes is an
equilibrium
state, regardless of the direction of the change. In contrast, an irreversible process is one in which
the
intermediate
states are not
equilibrium
9
states, so change occurs spontaneously in only one direction. As a result, a reversible process can
change direction at any time, whereas an irreversible process cannot. When a gas expands reversibly
against an external pressure such as a piston, for example, the expansion can be reversed at any time
by reversing the motion of the piston; once the gas is compressed, it can be allowed to expand again,
and the process can continue indefinitely. In contrast, the expansion of a gas into a vacuum
(Pext=0Pext=0) is irreversible because the external pressure is measurably less than the internal
pressure of the gas. No
equilibrium
states exist, and the gas expands irreversibly. When gas escapes from a microscopic hole in a balloon
into a vacuum, for example, the process is irreversible; the direction of airflow cannot change.
ΔSsysΔSsys for an Isothermal Expansion (or Compression)
As a substance becomes more dispersed in space, the thermal energy it carries is also spread over a
larger
volume
, leading to an increase in its entropy. Because entropy, like energy, is an extensive property, a dilute
solution of a given substance may well possess a smaller entropy than the same
volume
of a more concentrated solution, but the entropy per mole of solute (the molar entropy) will of
course always increase as the solution becomes more dilute.
For gaseous substances, the
volume
and pressure are respectively direct and inverse measures of
concentration
. For an
ideal gas
that expands at a constant temperature (meaning that it absorbs heat from the
surroundings
to compensate for the work it does during the expansion), the increase in entropy is given by
ΔS=
R
ln(V2V1)(13.4.6)(13.4.6)ΔS=
R
ln(V2V1)
Note: If the gas is allowed to cool during the expansion, the relation becomes more complicated and
will best be discussed in a more advanced course.
Because the pressure of an
ideal gas
is inversely proportional to its
volume
, i.e.,
P=n
R
TV(13.4.7)(13.4.7)P=n
R
TV
we can easily alter Equation 13.4.613.4.6 to express the entropy change associated with a change in
the pressure of an
ideal gas
10
:
ΔS=
R
ln(P1P2)(13.4.8)(13.4.8)ΔS=
R
ln(P1P2)
Also the
concentration
c=n/Vc=n/V for an
ideal gas
is proportional to pressure
P=c
R
T(13.4.9)(13.4.9)P=c
R
T
we can expressing the entropy change directly in concentrations, we have the similar relation
ΔS=
R
ln(c1c2)(13.4.10)(13.4.10)ΔS=
R
ln(c1c2)
Although these equations strictly apply only to perfect gases and cannot be used at all for liquids and
solids, it turns out that in a dilute solution, the solute can often be treated as a gas dispersed in the
volume
of the solution, so the last equation can actually give a fairly accurate value for the entropy of
dilution of a solution. We will see later that this has important consequences in determining the
equilibrium
concentrations in a
homogeneous
reaction mixture.
Temperature Entropy Diagram
Entropy change of a system is given by . During the reversible process, the energy
transfer as heat to the system from the surroundings is given by
(24.1)
11
Figure 24.1
Refer to figure 24.1. Here T and S are chosen as independent variables. The is the area under
the curve. The first law of thermodynamics gives . Also for a reversible process, we
can write,
and (24.2)
Therefore,
(24.3)
For a cyclic process, the above equation reduces to
(24.4)
For a cyclic process represents the net heat interaction which is equal to the net work done by
the system. Hence the area enclosed by a cycle on a T − S diagram represents the net work done by a
system. For a reversible adiabatic process, we know that
(24.5)
or,
(24.6)
or,
(24.7)
Hence a reversible adiabatic process is also called an isentropic process. On a T − S diagram, the
Carnot cycle can be represented as shown in Fig 24.1. The area under the curve 1-2 represents the
energy absorbed as heat by the system during the isothermal process. The area under the
curve 3-4 is the energy rejected as heat by the system. The shaded area represents the net
work done by the system.
We have already seen that the efficiency of a Carnot cycle operating between two thermal reservoirs
at temperatures T1and T2 is given by
(24.8)
12
This was derived assuming the working fluid to be an ideal gas. The advantage of T − S diagram can
be realized by a presentation of the Carnot cycle on the T − S diagram. Let the system change its
entropy from to during the isothermal expansion process 1-2. Then,
(24.9)
and,
(24.10)
and,
or,
(24.11)
This law was developed by the German chemist Walther Nernst between the years 1906 and 1912.
Alternate Statements of the 3rd Law of Thermodynamics
The Nernst statement of the third law of thermodynamics implies that it is not possible for a process
to bring the entropy of a given system to zero in a finite number of operations.
The American physical chemists Merle Randall and Gilbert Lewis stated this law differently: when the
entropy of each and every element (in their perfectly crystalline states) is taken as 0 at absolute zero
temperature, the entropy of every substance must have a positive, finite value. However, the entropy
at absolute zero can be equal to zero, as is the case when a perfect crystal is considered.
The Nernst-Simon statement of the 3rd law of thermodynamics can be written as: for a condensed
system undergoing an isothermal process that is reversible in nature, the associated entropy change
approaches zero as the associated temperature approaches zero.
Another implication of the third law of thermodynamics is: the exchange of energy between two
thermodynamic systems (whose composite constitutes an isolated system) is bounded.
13
−S(T∗)>0S(T)−S(T∗)>0 for any T>T∗>0T>T∗>0. Since the entropy is always finite, ∞>S(T)
−S(T∗)>0∞>S(T)−S(T∗)>0, so that
∞>limT∗→0[S(T)−S(T∗)] >0∞>limT∗→0[S(T)−S(T∗)] >0
and
∞>limT∗→0∫TT∗CPT dT>0∞>limT∗→0∫T∗TCPT dT>0
For temperatures in the neighborhood of zero, we can expand the heat capacity, to arbitrary
accuracy, as a Taylor series polynomial in TT:
CP(T)=CP(0)+(∂CP(0)∂T)PT+12(∂2CP(0)∂T2)PT2+…CP(T)=CP(0)+(∂CP(0)∂T)PT+12(∂2CP(0)∂T2)PT2+…
The inequalities become
∞>limT∗→0{CP(0)lnTT∗ +(∂CP(0)∂T)P(T−T∗)+14(∂2CP(0)∂T2)P(T−T∗)2+…} >0∞>lim
T∗→0{CP(0)lnTT∗ +(∂CP(0)∂T)P(T−T∗)+14(∂2CP(0)∂T2)P(T−T∗)2+…} >0
The condition on the left requires CP(0)=0CP(0)=0.
We could view the third law as a statement about the heat capacities of pure substances. We infer
not only that CP>0CP>0 for all T>0T>0, but also that
limT→0(CPT)=0 limT→0(CPT)=0
More generally, we can infer corresponding assertions for closed reversible systems that are not pure
substances: (∂H/∂T)P>0(∂H/∂T)P>0 for all T>0T>0,
and limT→0T−1(∂H/∂T)P=0 limT→0T−1(∂H/∂T)P=0 . (The zero-temperature entropies of such
systems are not zero, however.) In the discussion below, we describe the system as a pure substance.
We can make essentially the same arguments for any system; we need only
replace CPCP by (∂H/∂T)P(∂H/∂T)P. The Lewis and Randall statement asserts that the entropy goes to
a constant at absolute zero, irrespective of the values of any other thermodynamic functions. It
follows that the entropy at zero degrees is independent of the value of the pressure. For any two
pressures, P1P1 and P2P2, we have S(P2,0)−S(P1,0)=0S(P2,0)−S(P1,0)=0.
Letting P=P1P=P1 and P2=P+ΔPP2=P+ΔP and, we have
S(P+ΔP,0)−S(P,0)ΔP=0S(P+ΔP,0)−S(P,0)ΔP=0
for any ΔPΔP. Hence, we have
(∂S∂P)T=0=0
that is, T2−T1<t∗2−t1T2−T1<t2∗−t1>. Equivalently, the reversible process reaches a lower
temperature: T2<t∗2T2<t2∗>. From
dS=CPTdT−(∂V∂T)PdPdS=CPTdT−(∂V∂T)PdP
we can calculate the entropy changes for these processes. For the reversible process, we calculate
ΔSrev=S(P2,T2)−S(P1,T1)ΔSrev=S(P2,T2)−S(P1,T1)
To do so, we first calculate
(ΔS)T=S(P2,T1)−S(P1,T1)(ΔS)T=S(P2,T1)−S(P1,T1)
for the isothermal reversible transformation from state P1P1, T1T1 to the state specified
by P2P2 and T1T1. For this step, dTdT is zero, and so
(ΔS)T=∫P2P1(∂V∂T)PdP(ΔS)T=∫P1P2(∂V∂T)PdP
We then calculate
(ΔS)P=S(P2,T2)−S(P2,T1)(ΔS)P=S(P2,T2)−S(P2,T1)
for the isobaric reversible transformation from state P2P2, T1T1 to state P2P2, T2T2. For this
transformation, dPdP is zero, and
(ΔS)P=−∫T2T1CPTdT(ΔS)P=−∫T1T2CPTdT
Then,
ΔSrev=S(P2,T2)−S(P1,T1)=∫T2T1CPTdT−∫P2P1(∂V∂T)PdP=0ΔSrev=S(P2,T2)−S(P1,T1)=∫T1T2CPTdT−∫P1P
2(∂V∂T)PdP=0
Because ΔSrev=0ΔSrev=0, the reversible process is unique; that is, given P1P1, T1T1, and P2P2, the
final temperature of the system is determined. We find T2T2 from
15
∫T2T1CPTdT=∫P2P1(∂V∂T)PdP
What is Enthalpy?
Enthalpy is the measurement of energy in a thermodynamic system. The quantity of enthalpy equals
to the total content of heat of a system, equivalent to the system’s internal energy plus the product
of volume and pressure.
For more content on thermodynamics click here.
Technically, enthalpy describes the internal energy that is required to generate a system and the
amount of energy that is required to make room for it by establishing its pressure and volume and
displacing its environment.
When a process begins at constant pressure, the evolved heat (either absorbed or released) equals
the change in enthalpy. Enthalpy change is the sum of internal energy denoted by U and product of
volume and Pressure, denoted by PV, expressed in the following manner.
H=U+PV
Enthalpy is also described as a state function completely based on state functions P, T and U. It is
normally shown by the change in enthalpy (ΔH) of a process between the beginning and final states.
ΔH=ΔU+ΔPV
If the pressure and temperature don’t change throughout the process and the task is limited to
pressure and volume, the change in enthalpy is given by,
ΔH=ΔU+PΔV
The flow of heat (q) at constant pressure in a process equals the change in enthalpy based on the
following equation,
ΔH=q
Gibbs Free Energy
Gibbs free energy, also known as the Gibbs function, Gibbs energy, or free enthalpy, is a quantity
that is used to measure the maximum amount of work done in a thermodynamic system when the
temperature and pressure are kept constant. Gibbs free energy is denoted by the symbol ‘G’. Its
value is usually expressed in Joules or Kilojoules. Gibbs free energy can be defined as the maximum
amount of work that can be extracted from a closed system.
Download Complete Chapter Notes of Thermodynamics
Download Now
16
This property was determined by American scientist Josiah Willard Gibbs in the year 1876 when he
was conducting experiments to predict the behaviour of systems when combined together or
whether a process could occur simultaneously and spontaneously. Gibbs free energy was also
previously known as “available energy.” It can be visualised as the amount of useful energy present in
a thermodynamic system that can be utilised to perform some work.
Table of Contents
Gibbs Free Energy Equation
Second Law of Thermodynamics
Calculating Change in Gibbs Free Energy
Relationship between Free Energy and Equilibrium Constant
Relationship between Gibbs Free Energy and EMF of a Cell
Gibbs Free Energy Problems
Gibbs Free Energy Equation
Gibbs free energy is equal to the enthalpy of the system minus the product of the temperature
and entropy. The equation is given as:
G = H – TS
Where,
G = Gibbs free energy
An artificial neural network called an auto-encoder is used to encode data efficiently. The total cost
of the original code and the rebuilt code is calculated here using Helmholtz energy.
H = enthalpy
T = temperature
S = entropy
OR
or more completely as:
G = U + PV – TS
Where,
U = internal energy (SI unit: joule)
P = pressure (SI unit: pascal)
V = volume (SI unit: m3)
T = temperature (SI unit: kelvin)
S = entropy (SI unit: joule/kelvin)
Variations of the Equation
Gibbs free energy is a state function; hence it doesn’t depend on the path. So, change in Gibbs free
energy is equal to the change in enthalpy minus the product of temperature and entropy change of
the system.
ΔG = ΔH – Δ(TS)
If the reaction is carried out under constant temperature {ΔT=O}
ΔG = ΔH – TΔS
This equation is called the Gibbs-Helmholtz equation.
ΔG > 0; the reaction is non-spontaneous and endergonic
ΔG < 0; the reaction is spontaneous and exergonic
ΔG = 0; the reaction is at equilibrium
Note:
1. According to the second law of thermodynamics, the entropy of the universe always
increases in a spontaneous process.
2. ΔG determines the direction and extent of chemical change.
3. ∆G is meaningful only for reactions in which the temperature and pressure remain constant.
The system is usually open to the atmosphere (constant pressure), and we begin and end the
process at room temperature (after any heat that we have added or which is liberated by the
reaction has dissipated).
4. ∆G serves as the single master variable that determines whether a given chemical change is
thermodynamically possible. Thus, if the free energy of the reactants is greater than that of
the products, the entropy of the world will increase when the reaction takes place as written,
and so the reaction will tend to take place spontaneously. ΔS universe = ΔS system + ΔS
surroundings
5. If ΔG is negative, the process will occur spontaneously and is referred to as exergonic.
6. Therefore, spontaneity is dependent on the temperature of the system.
Modeling the dependence of the Gibbs and Helmholtz functions behave with varying temperature,
pressure, and volume is fundamentally useful. But in order to do that, a little bit more development
is necessary. To see the power and utility of these functions, it is useful to combine the First and
Second Laws into a single mathematical statement. In order to do that, one notes that since
dS=dqTdS=dqT
for a reversible change, it follows that
dq=TdSdq=TdS
And since
dw=TdS−pdVdw=TdS−pdV
for a reversible expansion in which only p-V works is done, it also follows that
(since dU=dq+dwdU=dq+dw):
dU=TdS−pdVdU=TdS−pdV
This is an extraordinarily powerful result. This differential for dUdU can be used to simplify the
differentials for HH, AA, and GG. But even more useful are the constraints it places on the variables T,
S, p, and V due to the mathematics of exact differentials!
Maxwell Relations
The above result suggests that the natural variables of internal energy are SS and VV (or the function
can be considered as U(S,V)U(S,V)). So the total differential (dUdU) can be expressed:
dU=(∂U∂S)VdS+(∂U∂V)SdVdU=(∂U∂S)VdS+(∂U∂V)SdV
Also, by inspection (comparing the two expressions for dUdU) it is apparent that:
(∂U∂S)V=T(22.3.1)(22.3.1)(∂U∂S)V=T
and
(∂U∂V)S=−p(22.3.2)(22.3.2)(∂U∂V)S=−p
But the value doesn’t stop there! Since dUdU is an exact differential, the Euler relation must hold
that
[∂∂V(∂U∂S)V]S=[∂∂S(∂U∂V)S]V[∂∂V(∂U∂S)V]S=[∂∂S(∂U∂V)S]V
By substituting Equations 22.3.122.3.1 and 22.3.222.3.2, we see that
[∂∂V(T)V]S=[∂∂S(−p)S]V[∂∂V(T)V]S=[∂∂S(−p)S]V
or
(∂T∂V)S=−(∂p∂S)V(∂T∂V)S=−(∂p∂S)V
This is an example of a Maxwell Relation. These are very powerful relationship that allows one to
substitute partial derivatives when one is more convenient (perhaps it can be expressed entirely in
terms of αα and/or κTκT for example.)
A similar result can be derived based on the definition of HH.
H≡U+pVH≡U+pV
Differentiating (and using the chain rule on d(pV)d(pV)) yields
dH=dU+pdV+VdpdH=dU+pdV+Vdp
Making the substitution using the combined first and second laws (dU=TdS–pdVdU=TdS–pdV) for a
reversible change involving on expansion (p-V) work
dH=TdS–pdV+pdV+VdpdH=TdS–pdV+pdV+Vdp
This expression can be simplified by canceling the pdVpdV terms.
dH=TdS+Vdp(22.3.3)(22.3.3)dH=TdS+Vdp
And much as in the case of internal energy, this suggests that the natural variables
of HH are SS and pp. Or
dH=(∂H∂S)pdS+(∂H∂p)SdV(22.3.4)(22.3.4)dH=(∂H∂S)pdS+(∂H∂p)SdV
Comparing Equations 22.3.322.3.3 and 22.3.422.3.4 show that
(∂H∂S)p=T(22.3.5)(22.3.5)(∂H∂S)p=T
and
19
(∂H∂p)S=V(22.3.6)(22.3.6)(∂H∂p)S=V
It is worth noting at this point that both (Equation 22.3.122.3.1)
(∂U∂S)V(∂U∂S)V
and (Equation 22.3.522.3.5)
(∂H∂S)p(∂H∂S)p
are equation to TT. So they are equation to each other
(∂U∂S)V=(∂H∂S)p(∂U∂S)V=(∂H∂S)p
Morevoer, the Euler Relation must also hold
[∂∂p(∂H∂S)p]S=[∂∂S(∂H∂p)S]p[∂∂p(∂H∂S)p]S=[∂∂S(∂H∂p)S]p
so
(∂T∂p)S=(∂V∂S)p(∂T∂p)S=(∂V∂S)p
This is the Maxwell relation on HH. Maxwell relations can also be developed based on AA and GG.
The results of those derivations are summarized in Table 6.2.1
Derivation of Maxwell’s law of distribution of velocities and
its experimental verification
Introduction
The kinetic molecular theory is used to determine the motion of a molecule of an ideal gas under a
certain set of conditions. However, when looking at a mole of ideal gas, it is impossible to measure
the velocity of each molecule at every instant of time. Therefore, the Maxwell-Boltzmann
distribution is used to determine how many molecules are moving between
velocities vv and v+dvv+dv. Assuming that the one-dimensional distributions are independent of one
another, that the velocity in the y and z directions does not affect the x velocity, for example, the
Maxwell-Boltzmann distribution is given by
dNN=(m2πkBT)1/2e−mv22kBTdv(3.1.2.1)(3.1.2.1)dNN=(m2πkBT)1/2e−mv22kBTdv
where
dN/NdN/N is the fraction of molecules moving at velocity vv to v+dvv+dv,
mm is the mass of the molecule,
kbkb is the Boltzmann constant, and
TT is the absolute temperature.1
Additionally, the function can be written in terms of the scalar quantity speed c instead of the vector
quantity velocity. This form of the function defines the distribution of the gas molecules moving at
different speeds, between c1c1 and c2c2, thus
f(c)=4πc2(m2πkBT)3/2e−mc22kBT(3.1.2.2)(3.1.2.2)f(c)=4πc2(m2πkBT)3/2e−mc22kBT
Finally, the Maxwell-Boltzmann distribution can be used to determine the distribution of the kinetic
energy of for a set of molecules. The distribution of the kinetic energy is identical to the distribution
of the speeds for a certain gas at any temperature. 2
Plotting the Maxwell-Boltzmann Distribution Function
Figure 1 shows the Maxwell-Boltzmann distribution of speeds for a certain gas at a certain
temperature, such as nitrogen at 298 K. The speed at the top of the curve is called the most probable
speed because the largest number of molecules have that speed.
20
will have a smaller speed distribution, while lighter molecules will have a speed distribution that is
more spread out.
Figure 3: The
speed probability density functions of the speeds of a few noble gases at a temperature of 298.15 K
(25 °C). The y-axis is in s/m so that the area under any section of the curve (which represents the
probability of the speed being in that range) is dimensionless. Figure is used with permission from
Wikipedia.
Related Speed Expressions
Three speed expressions can be derived from the Maxwell-Boltzmann distribution: the most
probable speed, the average speed, and the root-mean-square speed. The most probable speed is
the maximum value on the distribution plot. This is established by finding the velocity when the
following derivative is zero
df(c)dc|Cmp=0df(c)dc|Cmp=0
which is
Cmp=2RTM−−−−−√(3.1.2.3)(3.1.2.3)Cmp=2RTM
The average speed is the sum of the speeds of all the molecules divided by the number of molecules.
Cavg=∫∞0cf(c)dc=8RTπM−−−−−√(3.1.2.4)(3.1.2.4)Cavg=∫0∞cf(c)dc=8RTπM
The root-mean-square speed is square root of the average speed-squared.
Crms=3RTM−−−−−√(3.1.2.5)(3.1.2.5)Crms=3RTM
where
RR is the gas constant,
TT is the absolute temperature and
MM is the molar mass of the gas.
It always follows that for gases that follow the Maxwell-Boltzmann distribution (if thermalized)
Cmp<Cavg<Crms(3.1.2.6)(3.1.2.6)Cmp<Cavg<Crms
Concept of mean free path
In a gaseous system, the molecules never move in a straight path without interruptions. This is
because they collide with each other and change speed and direction. Between every two collisions,
a molecule travels a path length. The mean free path is the average of all path lengths between
collisions. In this article, we will derive an expression for the mean free path.
22
Table of Contents
What is Mean Free Path?
Derivation of Mean Free Path
Mean Free Path Factors
Frequently Asked Questions – FAQs
Source:en.wikipedia.org
Energy Density Formula
What is energy density?
23
Energy Density can be defined as the total amount of energy in a system per unit volume. For
example, the number of calories available per gram weight of the food. Foodstuffs have low energy
density and will provide less energy per gram of the food. It means that we can eat more of them
due to fewer calories.
Therefore we may say that energy density is the amount of energy accumulated in a system per unit
volume. It is denoted by letter U. Magnetic and electric fields are also the main sources for storing
the energy.
Energy Density Formula
In the case of electric field or capacitor, the energy density formula is expressed as below:
Electrical energy density = permittivity×Electricfieldsquared2In the form of equation,
UE = 12ε0E2
The energy density formula in case of magnetic field or inductor is as below:
Magnetic energy density = magneticfieldsquared2×magneticpermeability
In the form of an equation,
UB = 12μ0B2
The general energy is:
U = UE+UB
Where,
Regarding the electromagnetic waves, both magnetic and electric fields are involved in contributing
to energy density equally. Thus, the formula of energy density will be the sum of the energy density
of electric and magnetic fields both together.
Solved Examples
Q.1: In a certain region of space, the magnetic field has a value of 3×10−2 T. And the electric field has
a value of 9×107Vm−1. Determine the combined energy density of the electric and magnetic fields
both.
Solution: First we have to calculate the density and energy of each field separately. Then we will add
the densities to obtain the total energy density.
Given parameters in the question are:
B = 3×10−2T
E = 9×107Vm−1
ε0=8.85×10−12C2N−1m−2
μ=4×π×10−7NA−2 .
Thus electrical energy density,
UE = 12ε0E2
= 12×8.85×10−12×(9×107)2
= 12×8.85×10−12×(9×107)2
= 35842.5Jm−3
Now, magnetic energy density,
UB = 12μ0B2
= 12×4×π×10−7×(3×10−2)2
= 12×4×3.14×10−7×(3×10−2)2
= 358.1Jm−3
Thus total energy density will be,
U =UE+UB
= 35,842.5 + 358.1
= 36200.6 Jm−3
24
0,hv,2hv,3hv....nhv
If N0, N1,N2....are the number of oscillators per unit volume of the hologram possessing
energies0,hv,2hv....respectively, then the total number of oscillators N per unit volume will be
N = N0+N1+N2+.....
Nr = N0
= N0(1+ + +.....)
Hence the average energy per unit value (i.e., energy density) inside the enclosure is obtained by
multiplying with i.e. it is given by
Putting and ,the average energy per unit volume in the enclosure of the wave
a0=sinθπθπ.(2.11.12)(2.11.12)a0=sinθπθπ.
We have therefore arrived at the Fourier expansion of cosθxcosθx:
cosθx=2θsinθππ(12θ2−cosxθ2−12+cos2xθ2−22−cos3xθ2−32+...).(2.11.13)
(2.11.13)cosθx=2θsinθππ(12θ2−cosxθ2−12+cos2xθ2−22−cos3xθ2−32+...).
Put x=πx=π and rearrange slightly:
πcotθπ−1θ=2θ(1θ2−12+1θ2−22+...).(2.11.14)(2.11.14)πcotθπ−1θ=2θ(1θ2−12+1θ2−22+...).
Since we are assuming that θθ is some number between 0 and 1, we shall re-write this so that the
denominators are all positive:
πcotθπ−1θ=−2θ12−θ2−2θ22−θ2−...(2.11.15)(2.11.15)πcotθπ−1θ=−2θ12−θ2−2θ22−θ2−...
Now multiply both sides by dθdθ and integrate from θ=0θ=0 to θ=αθ=α. The integration must be
done with care. The indefinite integral of the left hand side
is lnsinθπ−lnθ+constantlnsinθπ−lnθ+constant, i.e. ln(sinθπθ)+constantln(sinθπθ)+constant. The
definite integral between 00 and αα is ln(sinαπα)−limθ→ 0ln(sinθπθ)ln(sinαπα)−limθ→ 0ln(sinθπθ).
The limit of the second term is lnπlnπ, so the definite integral is ln(sinαπαπ)ln(sinαπαπ). Integrating
the right hand side is a bit easier, so we arrive at
ln(sinαπαπ)=ln(12−α212)+ln(22−α222)+...(2.11.16)
(2.11.16)ln(sinαπαπ)=ln(12−α212)+ln(22−α222)+...
On taking the antilogarithm, we arrive at the required infinite product:
sinαπαπ=[1−α2][1−(12α)2][1−(13α)2]...(2.11.17)(2.11.17)sinαπαπ=[1−α2][1−(12α)2][1−(13α)2]...
Now expand this as a power series in α2α2:
sinαπαπ=1+()α2+()α4+()α6+...(2.11.18)(2.11.18)sinαπαπ=1+()α2+()α4+()α6+...
The first one is easy, but subsequent ones rapidly get more difficult, but you do have to get at least as
far as α4α4.
Now compare this expansion with the ordinary Maclaurin expansion:
sinαπαπ=1−π23!α2+π45!α4−...(2.11.19)(2.11.19)sinαπαπ=1−π23!α2+π45!α4−...
and we arrive at the correct expressions for the Riemann ζζ-functions. We then get for Stefan's law:
M=2π5k415h3c2T4=σT4,(2.11.20)(2.11.20)M=2π5k415h3c2T4=σT4,
where σ=5.6705×10−8 W m−2K−4.σ=5.6705×10−8 W m−2K−4.
Questions
Finally, now that you have struggled through Riemann’s zeta-function, let’s just make sure that you
have understood the really simple stuff, so here are a couple of easy questions – and you won’t have
to bother with zeta-functions.
1. By what factor should the temperature of a black body be increased so that
a) The integrated radiance (over all frequencies) is doubled?
b) The frequency at which its radiance is greatest is doubled?
c) The spectral radiance per unit wavelength interval at its wavelength of maximum spectral radiance
is doubled?
2. A block of shiny silver (absorptance = 0.23) has a bubble inside it of radius 2.2cm2.2cm, and it is
held at a temperature of 1200K1200K.
A block of dull black carbon (absorptance = 0.86) has a bubble inside it of radius 4.3cm4.3cm, and it
is held at a temperature of 2300K2300K,
Calculate the ratio
Integrated radiation energy density inside the carbon bubbleIntegrated radiation energy density
insdie the silver bubble.(2.10.7)(2.10.7)Integrated radiation energy density inside the carbon
bubbleIntegrated radiation energy density insdie the silver bubble.
Deriving the Rayleigh-Jeans Radiation Law
28
The Rayleigh-Jeans Radiation Law was a useful, but not completely successful attempt at establishing the
functional form of the spectra of thermal radiation. The energy density uνuν per unit frequency interval at a
frequency νν is, according to the The Rayleigh-Jeans Radiation,
uν=8πν2kTc2uν=8πν2kTc2
where kk is Boltzmann's constant, TT is the absolute temperature of the radiating body, and cc is the speed of
light in a vacuum.
This formula fits the empirical measurements for low frequencies, but fails increasingly for higher frequencies.
The failure of the formula to match the new data was called the ultraviolet catastrophe. The significance of this
inadequate so-called law is that it provides an asymptotic condition which other proposed formulas, such as
Planck's, need to satisfy. It gives a value to an otherwise arbitrary constant in Planck's thermal radiation
formula.
The Derivation
Consider a cube of edge length LL in which radiation is being reflected and re-reflected off its walls. Standing
waves occur for radiation of a wavelength λλ only if an integral number of half-wave cycles fit into an interval in
the cube. For radiation parallel to an edge of the cube this requires
Lλ/2=mLλ/2=m
where mm is an integer or, equivalently
λ=2Lmλ=2Lm
Between two end points there can be two standing waves, one for each polarization. In the following the
matter of polarization will be ignored until the end of the analysis and there the number of waves will be
doubled to take into account the matter of polarization.
Since the frequency νν is equal to c/λc/λ, where cc is the speed of light
ν=cm2Lν=cm2L
It is convenient to work with the quantity qq, known as the wavenumber, which is defined as
q=2πλq=2πλ
and hence
q=2πνcq=2πνc
In terms of the relationship for the cube,
q=2πm2L=π(mL)q=2πm2L=π(mL)
and hence
q2=π2(mL)2q2=π2(mL)2
Another convenient term is the radian frequency ω=2πνω=2πν. From this it follows that q=ω/cq=ω/c.
If mXmX, mYmY, mZmZ denote the integers for the three different directions in the cube then the condition for
a standing wave in the cube is that
q2=π2[(mXL)2+(mYL)2+(mZL)2]q2=π2[(mXL)2+(mYL)2+(mZL)2]
which reduces to
m2X+m2Y+m2Z=4L2ν2c2mX2+mY2+mZ2=4L2ν2c2
Now the problem is to find the number of nonnegative combinations of (mXmX, mYmY, mZmZ) that fit between
a sphere of radius RR and and one of radius R+dRR+dR. First the number of combinations ignoring the
nonnegativity requirement can be determined.
The volume of a spherical shell of inner radius RR and outer radius R+dRR+dR is given by
dV=4πR2dRdV=4πR2dR
If
R=m2X+m2Y+m2Z−−−−−−−−−−−−−√R=mX2+mY2+mZ2
then
R=4L2ν2c2−−−−−−√=2LνcR=4L2ν2c2=2Lνc
and hence
dR=2Ldνc.dR=2Ldνc.
This means that
dV=4π(2Lνc)2(2Lc)dν=32π(L3ν2c3)dν
Now the nonnegativity require for the combinations (mXmX, mYmY, mZmZ) must be taken into account. For the
two dimensional case the nonnegative combinations are approximately those in one quadrant of circle. The
approximation arises from the matter of the combinations on the boundaries of the nonnegative quadrant. For
29
the three dimensional case the nonnegative combinations constitute approximately one octant of the total.
Thus the number dNdN for the nonnegative combinations of (mXmX, mYmY, mZmZ) in this volume is equal
to (1/8)dV(1/8)dV and hence
dN=4πν2(L3c3)dνdN=4πν2(L3c3)dν
The average kinetic energy per degree of freedom is ½kT½kT, where kk is Boltzmann's constant. For harmonic
oscillators there is an equality between kinetic and potential energy so the average energy per degree of
freedom is kTkT. This means that the average radiation energy EE per unit frequency is given by
dEdν=kT(dNdν)=4πkT(L3c3)ν2dEdν=kT(dNdν)=4πkT(L3c3)ν2
and the average energy density, uνuν, is given by
duνdν=(1L3)(dEdν)=4πkTν2c3duνdν=(1L3)(dEdν)=4πkTν2c3
The previous only considered one direction of polarization for the radiation. If the two directions of
polarization are taken into account a factor of 2 must be included in the above formula; i.e.,
duνdν=8πkTν2c3duνdν=8πkTν2c3
This is the Raleigh-Jeans Law of Radiation and holds empirically as the frequency goes to zero.
Deriving the Wien's Displacement Law from Planck's Law
Wien's displacement law states that the blackbody radiation curve for different temperatures
peaks at a wavelength inversely proportional to the temperature. The shift of that peak is a direct
consequence of the Planck radiation law which describes the spectral brightness of black body
radiation as a function of wavelength at any given temperature. However it had been discovered
by Wilhelm Wien several years before Max Planck developed that more general equation, and
describes the entire shift of the spectrum of black body radiation toward shorter wavelengths as
temperature increases.
Derive Wien's displacement law from Planck's law. Proceed as follows:
ρ(ν,T)=2hν3c3(ehνkBT−1)(1)(1)ρ(ν,T)=2hν3c3(ehνkBT−1)
We need to evaluate the derivative of Equation 11 with respect to νν and set it equal to zero to
find the peak wavelength.
ddν{ρ(ν,T)}=ddν⎧⎩⎨⎪⎪⎪⎪⎪⎪2hν3c3(ehνkBT−1)⎫⎭⎬⎪⎪⎪⎪⎪⎪=0(2)
(2)ddν{ρ(ν,T)}=ddν{2hν3c3(ehνkBT−1)}=0
This can be solved via the quotient rule or product rule for differentiation. Selecting the latter for
convenience requires rewriting Equation 22 as a product:
ddν{ρ(ν,T)}=2hc3ddν{(ν3)(ehνkBT−1)−1}=0(3)(3)ddν{ρ(ν,T)}=2hc3ddν{(ν3)(ehνkBT−1)−1}=0
applying the product rule (and power rule and chain rule)
=2hc3[(3ν2)(ehνkBT−1)−1−(ν3)(ehνkBT−1)−2(hkBT)ehνkBT]=0(4)(4)=2hc3[(3ν2)(ehνkBT−1)−1−(ν3)
(ehνkBT−1)−2(hkBT)ehνkBT]=0
so this expression is zero when
(3ν2)(ehνkBT−1)−1=(ν3)(ehνkBT−1)−2(hkBT)ehνkBT(5)(5)(3ν2)(ehνkBT−1)−1=(ν3)
(ehνkBT−1)−2(hkBT)ehνkBT
or when simplified
3(ehνkBT−1)−(hvkBT)ehνkBT=0(6)(6)3(ehνkBT−1)−(hvkBT)ehνkBT=0
We can do a substitution u=hνkBTu=hνkBT and Equation 66 becomes
3(eu−1)−ueu=0(7)(7)3(eu−1)−ueu=0
Finding the solutions to this equation requires using Lambert's W-functions and results numerically
in
u=3+W(−3e−3)≈2.8214(8)(8)u=3+W(−3e−3)≈2.8214
so unsubstituting the uu variable
u=hνkBT≈2.8214(9)(9)u=hνkBT≈2.8214
or
ν≈2.8214kBhT≈(2.8214)(1.38×10−23J/K)6.63×10−34JsT≈(5.8×1010Hz/K)T(10)(11)(12)
(10)ν≈2.8214kBhT(11)≈(2.8214)(1.38×10−23J/K)6.63×10−34JsT(12)≈(5.8×1010Hz/K)T
30
The consequence is that the shape of the blackbody radiation function would shift proportionally
in frequency with temperature. When Max Planck later formulated the correct blackbody radiation
function it did not include Wien's constant explicitly. Rather, Planck's constant h was created and
introduced into his new formula. From Planck's constant h and the Boltzmann constant k, Wien's
constant (Equation 99) can be obtained.