You are on page 1of 62

1.

First Law of Thermodynamics


23 July 2021 19:30

Statement: The total energy of any system and its surroundings is conserved.

 The region in which the process occurs is set apart as the system.
 Everything with which the system interacts is its surroundings.

Δ(Energy of the system) + Δ(Energy of surroundings) = 0

 In the context of thermodynamics, heat and work represent energy in transit across the boundary dividing the system from its surroundings, and
are never stored or contained in the system. They represent the energy flow to or from a system.

 Potential, kinetic, and internal energy, on the other hand, reside with and are stored with matter.

• Energy balance on closed systems

If the boundary of a system does not permit the transfer of matter between the system and its surroundings, the system is said to be closed, and its mass
is necessarily constant.

Processes in which matter crosses the system boundary as streams that enter and leave process equipment. Such systems are said to be open.

Because no streams enter or leave a closed system, no energy associated with matter is transported across the boundary that divides the system from its
surroundings. All energy exchange between a closed system and its surroundings is in the form of heat or work, and the total energy change of the
surroundings equals the net energy transferred to or from it as heat and work.

Δ(Energy of surroundings) = ±Q ± W

 Sign Convention

Q and W will be positive for transferring into the system. So, the surroundings will have opposite sign.

Δ(Energy of surroundings) = Qsurr + Wsurr = െQ െ W


and,
Δ(Energy of the system) = Q + W

Closed systems often undergo processes during which only the internal energy of the system changes. For such processes,

ΔUt = Q + W

Total volume Vt and total internal energy Ut depend on the quantity of material in a system, and are called extensive properties. In contrast, temperature
and pressure, the principal thermodynamic coordinates for pure homogeneous fluids, are independent of the quantity of material, and are known as
intensive properties. For a homogeneous system, an alternative means of expression for the extensive properties, such as Vt and Ut , is:

Vt = mV or Vt = nV and Ut = mU or Ut = nU

New Section 1 Page 1


where the plain symbols V and U represent the volume and internal energy of a unit amount of material, either a unit mass or a mole. These are specific
or molar properties, respectively, and they are intensive, independent of the quantity of material actually present.

For a closed system,

For 1 mole, ΔU = Q + W or dU = dQ + dW

• Equilibrium and the Thermodynamic State

Equilibrium is a word denoting a static condition, the absence of change. In thermodynamics it means not only the absence of change but the absence of
any tendency toward change on a
macroscopic scale. Because any tendency toward change is caused by a driving force of one kind or another, the absence of such a tendency indicates
also the absence of any driving force. Hence, for a system at equilibrium all forces are in exact balance.

If this gas is heated or cooled, compressed or expanded, and then returned to its initial temperature and pressure, its intensive properties are restored to
their initial values. They do not depend on the past history of the substance nor on the means by which it reaches a given state. They depend only on
present conditions, however reached. Such quantities are known as state functions.

For a homogeneous pure substance, if two state functions are held at fixed values the thermodynamic state of the substance is fully determined. This
means that a state function, such as specific internal energy, is a property that always has a value; it may therefore be expressed mathematically as a
function of such coordinates as temperature and pressure, or temperature and density, and its values may be identified with points on a graph.

Heat and Work quantities, are not properties; they account for the energy changes that occur in the surroundings. They depend on the nature of the
process, and they may be associated with areas
rather than points on a graph.

The differential of a state function represents an infinitesimal change in its value. Integration of such a differential results in a finite difference between two
of its values, e.g.:

The differentials of heat and work are not changes, but are infinitesimal amounts. When integrated, these differentials give not finite changes, but finite
amounts. Thus,

For a closed system undergoing the same change in state by several processes, experiment shows that the amounts of heat and work required

New Section 1 Page 2


For a closed system undergoing the same change in state by several processes, experiment shows that the amounts of heat and work required
differ for different processes, but the sum Q + W is the same for all processes.

• Reversible Process

A process is reversible when its direction can be reversed at any point by an infinitesimal change in external conditions.

The reversible process is ideal in that it produces a best possible result; it yields the minimum work input required or maximum work output attainable from
a specified process.

Key features:

 Can be reversed at any point by an infinitesimal change in external conditions

 Is never more than minutely removed from equilibrium

 Traverses a succession of equilibrium states

 Is frictionless

 Is driven by forces whose imbalance is infinitesimal in magnitude

 Proceeds infinitely slowly

 ∙ When reversed, retraces its path, restoring the initial state of system and surroundings

Closed system reversible process

For 1 mole of homogenous fluid contained in a closed system, the energy balance can be written as:

dU = dQ + dW

The work for a mechanically reversible, closed-system process is given by, dW = െPdV. Substitution into the equation yields:

dU = dQ െ PdV

For a constant volume change of state,


dU = dQ
On integration,
∆U = Q

The internal energy change for a mechanically reversible, constant-volume, closed-system process equals the amount of heat transferred into the system.
For a constant-pressure change of state:

dU + PdV = d(U + PV) = dQ

New Section 1 Page 3


The group U + PV naturally arises here and in many other applications. This suggests the definition, for convenience, of this combination as a new
thermodynamic property. Thus, the mathematical (and only) definition of enthalpy is:

H ≡ U + PV

For a constant pressure change of state,


dH = dQ
On integration,
∆H = Q

• Heat Capacity

Gases, liquids, and solids have capacity for heat. The smaller the temperature change caused in a substance by the transfer of a given quantity of heat, the
greater its capacity.

Heat capacity at constant volume (CV)

The constant-volume heat capacity of a substance is defined as:

The parentheses and subscript V indicate that the derivative is taken with volume held constant; i.e., U is considered a function of T and V.

On integration,

or,

Heat capacity at constant pressure (CP)

It can be defined as

• Mass and Energy Balance in open systems

New Section 1 Page 4


• Mass and Energy Balance in open systems

Flow measurement

Open systems are characterized by flowing streams; there are 4 common measures of flow:

∙ Mass flow rate, ࢓ሶ ∙ Molar flow rate, ࢔ሶ ∙ Volumetric flow rate, q ∙ Velocity, u

The measures of flow are interrelated as:

where M is molar mass and A is the cross-sectional area for flow. Importantly, mass and molar flow rates relate to velocity:

Mass balance in an open system

From mass balance,

where, second term is:

or,

This form of mass balance equation is known as the continuity equation.

New Section 1 Page 5


This form of mass balance equation is known as the continuity equation.

At steady state, the accumulation term will be zero. So,

or,

Because specific volume is reciprocal of density,

General energy balance

Consider the following control volume with one entrance and one exit

Because energy, like mass, is conserved, the rate of change of energy within the control volume equals the net rate of energy transfer into the control
volume. Streams flowing into and out of the control volume have associated with them energy in its internal, potential, and kinetic forms, and all may
contribute to the energy change of the system.

The net work done on the system when all the entrance and exit sections are taken into account is െΔ[(PV)࢓ሶ]fs.Other forms of work like shaft work,
expansion and contraction of the control volume are represented by ࢃሶ. So, the above equation can also be written as:

New Section 1 Page 6


Combination of terms in accord with the definition of enthalpy, H = U + PV, leads to:

For example, throttling is an iso enthalpic process i.e. ∆H = 0.

New Section 1 Page 7


2. Second Law of Thermodynamics
26 July 2021 14:19

There exists a property called entropy S, which for systems at internal equilibrium is an intrinsic property, functionally related to the measurable state
variables that characterize the system. Differential changes in this property are given by the equation:

dSt = dQrev /T

where St is the system (rather than the molar) entropy.

The Second Law of Thermodynamics: The entropy change of any system and its surroundings, considered together, and resulting from any real
process, is positive, approaching zero when the process approaches reversibility. Mathematically,

ΔStotal ൒ 0

The second law affirms that every process proceeds in such a direction that the total entropy change associated with it is positive, the limiting value of
zero being attained only by a reversible process. No process is possible for which the total entropy decreases.

Now, if we integrate the above equation, we get

Let the temperatures of the reservoirs be TH and TC with TH > TC. Heat quantity Q, transferred from one reservoir to the other, is the same for both
reservoirs. However, QH and QC have opposite signs: positive for the heat added to one reservoir and negative for the heat extracted from the other.
Therefore QH = െQC, and the entropy changes of the reservoirs at TH and at TC are:

So,

Because the heat-transfer process is irreversible, so ΔStotal > 0, and therefore

 No process is possible which consists solely of the transfer of heat from one temperature level to a higher one.

Heat Engines and Heat Pumps

New Section 1 Page 8


(a) Heat engine (b) Heat pump or refrigerator

Following steps are involved in a carnot cycle,

• Step 1: A system at an initial temperature of a cold reservoir at TC undergoes a reversible adiabatic process that causes its temperature to rise to
that of a hot reservoir at TH.

• Step 2: The system maintains contact with the hot reservoir at TH and undergoes a reversible isothermal process during which heat QH is
absorbed from the hot reservoir.

• Step 3: The system undergoes a reversible adiabatic process in the opposite direction of Step 1 that brings its temperature back to that of the
cold reservoir at TC.

• Step 4: The system maintains contact with the reservoir at TC, and undergoes a reversible isothermal process in the opposite direction of Step 2
that returns it to its initial state with rejection of heat QC to the cold reservoir.

From first law of thermodynamics,

Because the engine inevitably operates in cycles, its properties over a cycle do not change. Therefore ΔU = 0, and W = െQH െ QC.

The entropy change of the surroundings equals the sum of the entropy changes of the reservoirs. Because the entropy change of the engine over a

New Section 1 Page 9


The entropy change of the surroundings equals the sum of the entropy changes of the reservoirs. Because the entropy change of the engine over a
cycle is zero, the total entropy change is that of the heat reservoirs. Therefore

or,

At W = 0,

or, if ∆Stotal = 0 then,

 It is impossible to construct an engine that, operating in a cycle, produces no effect (in system and surroundings) other than the
extraction of heat from a reservoir and the performance of an equivalent amount of work.

The second law does not prohibit the continuous production of work from heat, but it does place a limit on how much of the heat taken into a
cyclic process can be converted into work.

The thermal efficiency of a heat engine is defined as the ratio of the work produced to the heat supplied to the engine. With respect to the engine, the
work W is negative. Thus,

or,

 The efficiency of engine depends only on the temperature levels and not upon the working substance of the engine.

 For two given heat reservoirs no engine can have a thermal efficiency higher than that of a Carnot engine.

Steps involved:

New Section 1 Page 10


T - s graph for Carnot cycle

P - V graph for Carnot cycle

Because ΔSt is a property change, it is independent of path and is given by SBt െ SAt. If the fluid is changed from state A to state B by an irreversible
process, the entropy change is again ΔSt = SBt െ SAt , but experiment shows that this result is not given by ‫׬‬dQ/T evaluated for the irreversible process
itself, because the calculation of entropy changes by this integral must, in general, be along reversible paths.

The entropy change of a heat reservoir, however, is always given by Q/T, where Q is the quantity of heat transferred to or from the reservoir at
temperature T, whether the transfer is reversible or irreversible.

If a process is reversible and adiabatic, dQrev = 0 and dSt = 0 then, it is known as isoentropic process.

Entropy change for ideal gas

New Section 1 Page 11


For one mole or a unit mass of fluid undergoing a mechanically reversible process in a closed system, the first law becomes:

Although derived for a mechanically reversible process, this equation relates properties only, independent of the process causing the change of state,
and is therefore a general equation for the calculation of entropy changes in the ideal-gas state.

New Section 1 Page 12


3. Volumetric Properties of Pure Fluids
27 July 2021 19:13

Single phase region

P - V diagrams of pure substance


P - T diagram of pure substance

For the regions of the diagram where a single phase exists, there is a unique relation connecting P, V, and T. Expressed analytically, as f (P, V, T) = 0,
such a relation is known as a PVT equation of state.

An equation of state may be solved for any one of the three quantities P, V, or T, given values for other two. For example, if V is considered a function
of T and P, then V = V(T, P), and

If we divide both sides of the equation by V, we get

New Section 1 Page 13


or,

or,

These are weak functions of temperature and pressure.

Ideal gas and ideal gas state

A gas can behave like ideal gas at a very low pressure and very high temperature.

The subscript "ig" denotes properties for ideal gas state.

Property relations for ideal gas

Putting PV = RT in the enthalpy equation we get,

New Section 1 Page 14


Process calculations for ideal gas

• Isothermal processes

Here, dU = dH = 0. So, dQ = dW which is equal to,

and,

So,

• Isobaric processes

Here, dP = 0. So,

So,

• Isochoric processes

Here dV = 0 and dW = 0.So,

New Section 1 Page 15


• Adiabatic processes

Here, dQ = 0. So,

or,

or,

where,

and, work done will be

or,

Irreversible processes

The work of an irreversible process is usually calculated by a two-step procedure:

• W is determined for a mechanically reversible process that accomplishes the same change of state as the actual irreversible process.

• Then, this result is multiplied or divided by an efficiency to give the actual work. If the process produces work, the absolute value for the reversible
process is larger than the value for the actual irreversible process and must be multiplied by an efficiency. If the process requires work, the value

New Section 1 Page 16


process is larger than the value for the actual irreversible process and must be multiplied by an efficiency. If the process requires work, the value
for the reversible process is smaller than the value for the actual irreversible process and must be divided by an efficiency.

Virial Equation of State

2 forms of Virial Equation of State

This dimensionless ratio is called the compressibility factor. It is a measure of the deviation of the real-gas molar volume from its ideal-gas value. For
the ideal-gas state, Z = 1. At moderate temperatures its value is usually <1, though at elevated temperatures it may be >1.

So,

or,

Both of these equations are known as virial expansions, and the parameters B′, C′, D′, etc., and B, C, D, etc., are called virial coefficients. Parameters
B′ and B are second virial coefficients; C′ and C are third virial coefficients, and so on.

Applications of Virial EOS

If we differentiate (1), we get

From which,

or,

New Section 1 Page 17


or,

Van der Waal Equation of State

This equation can be solved for the 2 constants by using 2 boundary conditions as:

which gives

and,

where, Ψ = 27/64 and Ω = 1/8. On solving the above equation using these values we can get critical compressibility factor (ZC) = 0.375

New Section 1 Page 18


PV isotherm given by cubic equation of state

Corresponding States

2 parameter theorem: All fluids, when compared at the same reduced temperature and reduced pressure, have approximately the same
compressibility factor, and all deviate from ideal-gas behaviour to about the same degree.

These correlations of Z require the use of only two reducing parameters Tc and Pc.

3 parameter theorem: All fluids having the same value of acentric factor, when compared at the same reduced temperature and reduced pressure,
have about the same compressibility factor, and all deviate from ideal-gas behaviour to about the same degree.

New Section 1 Page 19


Approximate temperature dependence of the reduced vapour pressure

The acentric factor for a pure chemical species is defined with reference to its vapor pressure. The logarithm of the vapor pressure of a pure fluid is
approximately linear in the reciprocal of absolute temperature. This linearity can be expressed by the Pitzer's correlation as:

Here, Pr = P/Pc and Tr = T/Tc

New Section 1 Page 20


4. Thermodynamic Properties of Fluids
28 July 2021 10:21

Fundamental Property Relations

These relations are applicable to any process in a closed PVT system resulting in a differential change from one equilibrium state to another.

Following are the fundamental relations (for 1 mole of substance):


dQrev = dU + dWrev

Putting dWrev = െ P dV and dQrev = T dS in the above equation

we have, dU = T dS െ P dV

dH = T dS + V dP

dA = െS dT െ P dV

dG = െS dT + V dP

Implicit in each of these equations is a functional relationship that expresses a molar (or unit mass) property as a function of a natural or special pair of
independent variables:

U = U(S, V) H = H(S, P) A = A(T, V) G = G(T, P)

These equation can also be written in the general form as:

or,
dF = M dx + N dy
where,

and,

then,

New Section 1 Page 21


and,

or,

So, for each of these exact differentials, we can write the relations like above which are known as Maxwell's Relations as shown below:

Enthalpy and Entropy relations as function of T & P

where,

also,

and,

New Section 1 Page 22


now,

so,

now, dH = T dS + VdP, differentiating this equation w.r.t. P, we get

Now, from maxwell's relations we get,

So,

For the case of ideal gases, these equations will become,

And for liquids,

Internal energy as function of P

Since we know that U = H െ PV. On differentiating this equation w.r.t. P we get,

New Section 1 Page 23


Since we know that U = H െ PV. On differentiating this equation w.r.t. P we get,

On putting,

Putting the values of the differentials on the R.H.S. from the volume expansivity and isothermal compressibility equations, we get

Internal energy and entropy as function of V and T

The partial derivative of U from dU = T dS - PdV,

or,

and,

So,

New Section 1 Page 24


and,

If we put,

then,

and,

Gibbs free energy as generating function

The fundamental property relation for G = G(T, P),


dG = V dP െ S dT

has an alternative form. It follows from the mathematical identity:

Now, if we put G = H - TS and dG = VdP - SdT in the above equation, we get

or,

and,

New Section 1 Page 25


also,

Residual Properties or Departure Functions

MR ≡ M െ Mig
For example,

Because, V = ZRT/P, the residual volume and the compressibility factor can be related as:

In the similar fashion, we can also write the generating function for the ideal gas,

If we subtract this from the generating function we get,

or,

and,

At constant temperature,

New Section 1 Page 26


If we put the value of VR in this equation and integrate it from 0 to P, we get,

ࡳ ࡾ ࡼ ࢊࡼ
ࡾࢀ = ‫׬‬૙ (ࢆ − ૚) ۛۛ
ۛۛ ࡼ (const T)
so,

and,

2 phase systems

The curves shown on the PT diagram represent phase boundaries for a pure substance. A phase transition at constant temperature and pressure
occurs whenever one of these curves is crossed, and as a result the molar or specific values of the extensive thermodynamic properties change
abruptly. Thus the molar or specific volume of a saturated liquid is very different from the molar or specific volume of saturated vapor at the same T
and P. This is true as well for internal energy, enthalpy, and entropy. The exception is the molar or specific Gibbs energy, which for a pure species
does not change i.e. dG = 0 during a phase transition such as melting, vaporization, or sublimation.

Temperature dependence of vapour pressure of liquids

New Section 1 Page 27


where A and B are constants for a given species. This equation gives a rough approximation of the vapor-pressure relation for the entire temperature
range from the triple point to the critical point.

A more accurate form of the equation is:

2 phase vapour - liquid system

When a system consists of saturated-liquid and saturated-vapor phases coexisting in equilibrium, the total value of any extensive property of the two-
phase system is the sum of the total properties of the phases. Written for the volume, this relation is:

where V is the molar volume for a system containing a total number of moles n = nl + nv. Division by n gives:

Putting xl = 1 - xv we get,

In this equation the properties V, Vl , and Vv may be either molar or unit-mass values. The mass or molar fraction of the system that is vapor xv is
often called the quality, particularly when the fluid in question is water. Analogous equations can be written for the other extensive thermodynamic
properties. All of these relations are represented by the generic equation

Alternatively,

where, M represents H, S, U, V etc.

Thermodynamic Diagrams

A thermodynamic diagram is a graph showing a set of properties for a particular substance, e.g., T, P, V, H, and S. The most common diagrams are:
TS, PH (usually ln P vs. H), and HS (called a Mollier diagram). The designations refer to the variables chosen for the coordinates.

The Mollier diagram does not usually include volume data. In the vapor or gas region, lines for constant temperature and constant superheat appear.
Superheat is a term denoting the difference between the actual temperature and the saturation temperature at the same pressure.

New Section 1 Page 28


Superheat is a term denoting the difference between the actual temperature and the saturation temperature at the same pressure.

Some process paths are easily traced on particular thermodynamic diagrams. For example, the boiler of a steam power plant has liquid water as feed
at a temperature below its boiling point, and superheated steam as product. Thus, water is heated at constant P to its saturation temperature (line
1–2), vaporized at constant T and P (line 2–3), and superheated at constant P (line 3–4).

Pressure - Enthalpy Graph Temperature - Entropy Graph

New Section 1 Page 29


5. Solution Thermodynamics
28 July 2021 17:37

Fundamental Property Relations

We know that,

d(nG) = (nV)dP െ (nS)dT

where n is the total number of moles of the system. It applies to a single-phase fluid in a closed system wherein no chemical reactions occur. For such a
system the composition is necessarily constant, and therefore:

The subscript n indicates that the numbers of moles of all chemical species are held constant.

For the more general case of a single-phase, open system, material may pass into and out of the system, and nG becomes a function of the numbers of
moles of the chemical species present. It remains a function of T and P, and we can therefore write the functional relation:

nG = g(T, P, n1, n2, …., ni,…)

where, ni is the no. of moles of species i. The total differential of nG is then,

The summation is over all species present, and subscript nj indicates that all mole numbers except the i th are held constant. The derivative in the final
term is given its own symbol and name. Thus, by definition the chemical potential of species i in the mixture is:

With this definition and with the first two partial derivatives replaced by (nV) and െ(nS), the preceding equation becomes:

This equation is the fundamental property relation for single - phase fluid systems of variable mass and composition and it is the foundation equation for
the solution thermodynamics. For the special case of n = 1 and ni = xi :

New Section 1 Page 30


Here, if the composition becomes constant, the equation will reduce to:

dG = VdP - SdT

Implicit in this equation is the functional relationship of the molar Gibbs energy to its canonical variables, here: T, P, and {xi}:

G = G(T, P, n1, n2, …., ni,…)


Also,

Chemical Potential and Equilibrium

If we equate d(nG) = (nV) dP - nS dT and d(nG) = (nV) dP - (nS)dT + ∑࢏ ࣆ࢏ ࢊ࢔࢏ we get,

With respect to phase equilibrium, we note that for a closed nonreacting system consisting of two phases in equilibrium, each individual phase is open to
the other, and mass transfer between phases may occur.

where, α and β identify the phases. For the system to be in thermal and mechanical equilibrium T and P must be uniform.

The change in the total Gibbs energy of the two-phase system is the sum of the equations for the separate phases. When each total-system property is
expressed by an equation of the form,

Comparison of the two equations shows that at equilibrium:

New Section 1 Page 31


The changes dni α and dni β result from mass transfer between the phases; mass conservation therefore requires:

Quantities dniα are independent and arbitrary, and the only way the left side of the second equation can in general be zero is for each term in parentheses
separately to be zero. Hence,

where N is the number of species present in the system. Successive application of this result to pairs of phases permits its generalization to multiple
phases; for π phases:

where, i = 1, 2, 3, …., N

Thus, multiple phases at the same T and P are in equilibrium when the chemical potential of each species is the same in all phases.

Partial Properties

The definition of the chemical potential as the mole-number derivative of nG suggests that other derivatives of this kind may prove useful in solution
thermodynamics. Thus, we define the partial molar property തതതത
ࡹ࢏ of species i in solution as:

Sometimes called a response function, it is a measure of the response of total property nM to the addition of an infinitesimal amount of species i to a
finite amount of solution, at constant T and P.

Molar (or unit-mass) properties are represented by the plain symbol M. Partial properties are denoted by an overbar, with a subscript to identify the
species; the symbol is therefore തതതത
ࡹ࢏ . In addition, properties of the individual species as they exist in the pure state at T and P of the solution are identified
only by script and the symbol is Mi. In summary the three kinds of properties used in thermodynamics are:

New Section 1 Page 32


So,

Relation between molar and partial molar properties

Earlier we have defined the partial molar properties as the differential of the molar properties. Now, we are calculating the molar properties from the partial
molar properties. Thus, for a property M, we can write nM as a function which we would call as ॸ:

The total differential of nM is:

where subscript n indicates that all mole numbers are held constant, and subscript nj that all mole numbers except ni, are kept constant. So,

where subscript x denotes differentiation at constant composition. Because ni = nxi,

Moreover,
d(nM) = ndM + Mdn

So, equating this value of d(nM) with the above equation of d(nM) we get,

The terms containing n are collected and separated from those containing dn to yield:

In application, one is free to choose a system of any size, as represented by n, and to choose any variation in its size, as represented by dn. Thus n and

New Section 1 Page 33


In application, one is free to choose a system of any size, as represented by n, and to choose any variation in its size, as represented by dn. Thus n and
dn are independent and arbitrary. The only way that the left side of this equation can then, in general, be zero is for each term in brackets to be zero.
Therefore,

and,
These relations are called "Summability Relations"
(they allow the calculation of mixture properties
Multiplying above equation by 'n' yields the alternative expression: from partial properties)

If we differentiate (1), we will get,

Combining this equation with the above equation of dM we will obtain the Gibbs - Duhem equation as:

This equation must be satisfied for all changes occurring in a homogeneous phase. Now, at constant T and P, it reduces to:

So, this shows that the partial molar properties cannot all vary independently. This constraint is analogous to the constraint on mole fractions, which
are not all independent because they must sum to one. Similarly, the mole-fraction-weighted sum of the partial molar properties (equation (1)) must yield
the overall solution property, and this constrains the variation in partial molar properties with composition.

Limiting cases for partial properties

ഥ ࢏ approach the
Partial properties, like solution properties, are functions of composition. In the limit as a solution becomes pure in species i, both M and ࡹ
pure-species property Mi. Mathematically,

For a partial property of a species that approaches its infinite-dilution limit, i.e., a partial property value of a species as its mole fraction approaches
zero, we can make no general statements. Values come from experiment or from models of solution behaviour. Because it is an important quantity, we do
give it a symbol, and by definition we write:

New Section 1 Page 34


Partial properties in binary solutions

Since we know that,

So, for binary mixtures we can write,

On differentiation, we get

When M is known as the function of x1 at constant T and P, the appropriate form of the Gibbs/Duhem equation is expressed here as:

Because x1 + x2 = 1, it follows that dx1 = െdx2. Eliminating dx2 in favour of dx1 in Eq. (2) and combining the result with Eq. (3) gives:

Two equivalent forms of Eq. (1) result from the elimination separately of x1 and x2 :

Combining this with (4), we get

and,

Equation (3), the Gibbs - Duhem Equation can also be written in the derivative form as:

and,

New Section 1 Page 35


and,

ഥ ૚ and ࡹ
Thus, plots of ࡹ ഥ ૛ vs. x1 become horizontal as each species approaches purity.

Graphically,

(a) M vs. x1 plot (b) Values of Partial properties at infinite dilution

Note: Here, I1 and I2 are given by:

New Section 1 Page 36


Note: Here, I1 and I2 are given by:

and,

Relations among partial properties

We know that,

and,

So,

And, from Maxwell's equation,

and,

or,

and,

These equations allow us to calculate the effects of P and T on the partial Gibbs energy (or chemical potential). They are the partial-property analogus of
the equations derived from the Gibbs free energy equation (d(nG) = (nV)dP െ (nS)dT) . Many additional relationships among partial properties can be
derived in the same ways that relationships among pure species properties were derived in earlier chapters. More generally, one can prove the following:

Every equation that provides a linear relation among thermodynamic properties of a constant-composition solution has as its counterpart an
equation connecting the corresponding partial properties of each species in the solution.

New Section 1 Page 37


By the definition of partial properties, this becomes:

or,

where,

and,

The Ideal Gas State Mixture Model

For ideal gases, the partial molar volume can be written as:

This result means that for the ideal-gas state at given T and P the partial molar volume, the pure-species molar volume, and the mixture molar volume are
identical:

New Section 1 Page 38


We define the partial pressure of species i in the ideal-gas-state mixture (pi) as the pressure that species i would exert if it alone occupied the molar
volume of the mixture. Thus,

where yi is the mole fraction of species i. The partial pressures obviously sum to the total pressure. Because the ideal-gas-state mixture model presumes
molecules of zero volume that do not interact, the thermodynamic properties (other than molar volume) of the constituent species are independent of one
another, and each species has its own set of private properties. This is the basis for the following statement of Gibbs’s theorem as:

A partial molar property (other than volume) of a constituent species in an ideal-gas-state mixture is equal to the corresponding molar property
of the species in the pure ideal-gas state at the mixture temperature but at a pressure equal to its partial pressure in the mixture.

This is expressed mathematically for generic partial property at

Enthalpy in the ideal-gas state is independent of pressure; therefore

More simply,

where Hiig is the pure-species value at the mixture T. An analogous equation applies for Uig and other properties that are independent of pressure.

Entropy in the ideal-gas state does depend on pressure, restricted to constant temperature:

This provides the basis for computing the entropy difference between a gas at its partial pressure in the mixture and at the total pressure of the mixture.
Integration from pi to P gives:

Hence,

New Section 1 Page 39


Hence,

If we use following expression,

we get,

where, Siig is the pure species value in the mixture and T and P.

For the Gibbs energy in the ideal-gas-state mixture,

The parallel relation for partial properties can be can be written as:

Putting the values of variables from (1) and (2) on the RHS we get,

or,

On applying the summability rule on (1), (2) and (3) we get,

Now, again we can rewrite these equations as:

New Section 1 Page 40


Now, in the following equation,

If we put G = Giig in dG = -SdT + VdP at constt. T (dT = 0) we get,

On integration, we get

where, Гi (T) is integration constant at constant T, is a species-dependent function of temperature only.

Also,

Now, from summability rule the Gibbs energy for the ideal-gas-state mixture is:

Fugacity and fugacity Coefficient: Pure Species

The chemical potential μi provides the fundamental criterion for phase equilibrium. This is true as well for chemical-reaction equilibria. However, it exhibits
characteristics that discourage its direct use. The Gibbs energy, and hence μi, is defined in relation to internal energy and entropy. Because absolute
values of internal energy are unknown, the same is true for μi.

Also, μiig approaches negative infinity when either P or yi approaches zero. This is true not only for the ideal-gas state, but for any gas. Although these
characteristics do not preclude the use of chemical potentials, the application of equilibrium criteria is facilitated by the introduction of the fugacity, a
property that takes the place of μi but does not exhibit its less desirable characteristics.

The word fugacity is based on a Latin root meaning to flee or escape, also the basis for the word fugitive. Thus fugacity has been interpreted to
mean “escaping tendency.” When the escaping tendency is the same for the two phases, they are in equilibrium. When the escaping tendency
of a species is higher in one phase than another, that species will tend to transfer to the phase where its fugacity is lower.

The origin of the fugacity concept resides in the following equation:

New Section 1 Page 41


The origin of the fugacity concept resides in the following equation:

which is valid only for pure species i in the ideal-gas state. For a real fluid, we write an analogous equation that defines fi, the fugacity of pure species i:

This new property fi, with units of pressure, replaces P in the above equation. If we take this new equation as a special case of the previous equation we
can say that:

And, if we subtract these 2 equations, we get:

or,

or,

since,

Here, the dimensionless ratio fi/P has been defined as another new property, the fugacity coefficient, given by symbol ϕi:

And, for ideal gas behaviour,

In the following equation,

If we put Zi - 1 = BiiP/RT (second Virial coefficient), we get

New Section 1 Page 42


If we put Zi - 1 = BiiP/RT (second Virial coefficient), we get

or,

VLE for pure species

For both phases, we can write

or,

At equilibrium, Giv - Gli = 0. So,

where, fisat indicates the value of either vapour or liquid. So,

For a pure species, coexisting liquid and vapor phases are in equilibrium when they have the same temperature, pressure, and fugacity.

Alternatively,

Fugacity of Pure Liquid (Poynting factor)

Fugacity and fugacity Coefficient: Species in Solution

The definition of the fugacity of a species in solution is parallel to the definition of the pure-species fugacity. For species i in a mixture of real gases or in a
solution of liquids, the equation analogous to partial molar Gibb's Free Energy , the ideal-gas-state expression is:

New Section 1 Page 43


where, ࢌ෠࢏ is the fugacity of species i in solution, replacing the partial pressure yiP. This definition of ࢌ෠࢏ does not make it a partial molar property.

A direct application of this definition indicates its potential utility. The equality of μi in every phase, is the fundamental criterion for phase equilibrium.
Because all phases in equilibrium are at the same temperature, an alternative and equally general criterion follows immediately from the above equation:

Thus, multiple phases at the same T and P are in equilibrium when the fugacity of each constituent species is the same in all phases.

Since we know that,

where M is the molar (or unit-mass) value of a thermodynamic property and Mig is the value that the property would have in its ideal-gas state with the
same composition at the same T and P. The defining equation for a partial residual property ࡹ࢏തതതതࡾ
 follows from this equation. Multiplied by n mol of
mixture, it becomes:

This is in the form of the partial molar property

In the same way, writing for the residual Gibb's energy, we have,

or,

So,

where,

which is known as the fugacity coefficient of the species 'i' in the solution. Although most commonly applied to gases, the fugacity coefficient may also

New Section 1 Page 44


which is known as the fugacity coefficient of the species 'i' in the solution. Although most commonly applied to gases, the fugacity coefficient may also
be used for liquids, and in this case mole fraction yi is replaced by xi, the symbol traditionally used for mole fractions in the liquid phase.

Thus, the fugacity of species 'i' in an ideal-gas-state mixture is equal to its partial pressure. Moreover, for ideal gases:

Fundamental Residual Property Relations

For 'n' moles, we know that,

Putting G = H - TS and eliminating nG by:

We get,

For ideal gases also,

On subtracting these 2 equations, we will get,

or,

and,

New Section 1 Page 45


This equation demonstrates that the logarithm of the fugacity coefficient of a species in solution is a partial property with respect to GR/RT.

Ideal Solution Model

Like earlier, we have defined the chemical potential for the ideal gas state mixture model. In the same fashion, we will define the chemical potential for the
ideal solution model also as:

where, superscript id denotes an ideal-solution property. Mole fraction is here represented by xi to reflect the fact that this approach is most often applied
to liquids.

All other thermodynamic properties for an ideal solution follow from the above equation. The partial volume results from differentiation with respect to
pressure at constant temperature and composition gives:

or,

similarly,

or,

now,

because,

or,

Now, applying the summability rule on these equations, we get

New Section 1 Page 46


Now, applying the summability rule on these equations, we get

Lewis/Randall Rule

We already know that,

On subtracting these 2 equations we will get,

And, for ideal solutions we know that,

If we compare the above equation with the equation given below,

we get,

This equation is known as Lewis/Randall rule applies to each species in an ideal solution at all conditions of temperature, pressure, and composition. It
shows that the fugacity of each species in an ideal solution is proportional to its mole fraction; the proportionality constant is the fugacity of pure
species i in the same physical state as the solution and at the same T and P.

On dividing both sides by Pxi and putting

New Section 1 Page 47


we will get,

Thus the fugacity coefficient of species i in an ideal solution is equal to the fugacity coefficient of pure species i in the same physical state as the solution
and at the same T and P.

Excess Properties

Similar to the residual properties for the gases, there exist excess properties for liquids which can be defined as the difference between the actual
property value of a solution and the value it would have as an ideal solution at the same temperature, pressure, and composition. Thus:

where, M can be H, U, V, S, G, etc.

For example, SE ≡ S - Sid HE ≡ H - Hid GE ≡ G - Gid

Also, GE = HE - TSE

Fundamental excess property relation:

 Excess properties are often strong functions of temperature, but at normal temperatures they are not strongly influenced by pressure.

Property Changes in Mixing

Total volume change of the system after mixing,

Because the process occurs at constant pressure, the heat transfer Q is equal to the total enthalpy change of the system:

Dividing these equations by n1 + n2 , we get:

and,

New Section 1 Page 48


Thus the volume change of mixing ΔV and the enthalpy change of mixing ΔH are found from the measured quantities ΔVt and Q. Because of its
experimental association with Q, ΔH is usually called the heat of mixing.

Equations analogous to those for ΔV and ΔH can be written for any property, and they may also be generalized to apply to the mixing of any number of
species:

where M can represent any intensive thermodynamic property of the mixture, for example, U, CP, S, G, or Z. This equation is known as the property
change of mixing.

Property changes of mixing are functions of T, P and composition and are directly related to excess properties. So,

or,

or,
∆H = H - x1H1 - x2H2

For the special case of an ideal solution, each excess property is zero, and these equations become:

∆Vid = 0 ∆Hid = 0

W = T∆S = -RT∑ ࢞࢏ ࢒࢔࢞࢏


or for n moles,
W = T∆S = -nRT  ࢞࢏ ࢒࢔࢞࢏

New Section 1 Page 49


W = T∆S = -nRT∑ ࢞࢏ ࢒࢔࢞࢏

For a binary system the partial excess properties are given by:

Putting M = ME we have,

Here, x2 = 1 - x1

New Section 1 Page 50


6. Vapour - Liquid Equilibrium Thermodynamics
28 July 2021 21:08

Equilibrium is a condition in which no changes occur in the macroscopic properties of an isolated system with time. At equilibrium, all potentials that may
cause change are exactly balanced, so no driving force exists for any change in the system. An isolated system consisting of liquid and vapor phases in
intimate contact eventually reaches a final state wherein no tendency exists for change to occur within the system. The temperature, pressure, and phase
compositions reach final values which thereafter remain fixed.

Nevertheless, at the microscopic level, conditions are not static. The molecules comprising a phase at a given instant are not the same molecules that later
occupy the same phase. Molecules constantly pass from one phase to the other. However, the average rate of passage of molecules is the same in both
directions, and no net interphase transfer of material occurs.

Excess Gibbs Energy & Activity Coefficients

We know that,

And for an ideal solution,

By difference,

Above equation is written for the partial Gibbs energy, shows that the left side of this equation is the partial excess Gibbs energy ; the dimensionless ratio
appearing on the right is the activity coefficient of species i in solution, symbol ࢽ࢏ . Thus, by definition,

Whence,

So, the following equation

can be re- written as:

Also,

New Section 1 Page 51


And,

We know that ln ࢽ࢏ is a partial property w.r.t GE /RT :

At constant T and P,

This is known as Gibbs - Duhem Equation

For binary mixture, these equations can be written as:

And,

Modified Raoult's Law

For azeotropic mixtures, yi = xi

The criterion for vapor/liquid equilibrium is that the fugacity of species i is the same in both phases. If the vapor phase is in its ideal-gas state, then the
fugacity equals the partial pressure, and

New Section 1 Page 52


The liquid-phase fugacity of species i increases from zero at infinite dilution (xi = yi → 0) to Pisat for pure species i.

Correlations for Liquid - phase Activity Coefficents

GEോRT is a function of T, P and composition, but for liquids at low to moderate pressures it is a very weak function of P. Therefore its pressure dependence
is usually neglected, and for applications at constant T, excess Gibbs energy is treated as a function of composition alone:

Margules Equation

For the limiting conditions of infinite dilution, they imply:

Van Laar Equation

New Section 1 Page 53


7. Heat Effects
30 July 2021 21:18

Sensible Heat Effects

When the system is a homogeneous substance of constant composition, the phase rule indicates that fixing the values of two intensive properties
establishes its state. The molar or specific internal energy of a substance may therefore be expressed as a function of two other state variables. The
key thermodynamic variable is temperature. With molar or specific volume chosen arbitrarily, we have U = U(T, V). Then,

or,

The second term on RHS will be zero in two circumstances:

• For any closed-system constant-volume process.

• Whenever the internal energy is independent of volume, as for the ideal-gas state and the incompressible liquid.

In either cases,

and,

Although real liquids are to some degree compressible, far below their critical temperature they can often be treated as incompressible fluids. The
ideal-gas state is also of interest, because actual gases at low pressures approach ideality. The only possible mechanically reversible constant-
volume process is simple heating (stirring work is inherently irreversible), for which Q = ΔU which is given as:

The enthalpy may be treated similarly, with molar or specific enthalpy expressed most conveniently as a function of temperature and pressure. Then
H = H(T, P), and

New Section 1 Page 54


or,

Again, the second term on RHS is zero for two situations:

• For any constant-pressure process.

• When the enthalpy is independent of pressure, regardless of process. This is exactly true for the ideal-gas state and approximately true for real
gases at low pressure and high temperature.

In either cases,

and,

Moreover, Q = ΔH for mechanically reversible, constant-pressure, closed-system processes and WS = 0. In either case,

Latent Heat of Pure Substances

When a pure substance is liquefied from the solid state or vaporized from the liquid or solid at constant pressure, no change in temperature occurs;
however, these processes require the transfer of finite amounts of heat to the substance. These heat effects are called latent heats: of fusion, of
vaporization, and of sublimation.

The characteristic feature of all these processes is the coexistence of two phases. According to the phase rule, the intensive state of a two-phase
system consisting of a single species is fixed by specification of just one intensive property. Thus the latent heat accompanying a phase change is a
function of temperature only, and is related to other system properties by an exact thermodynamic equation:

Clapeyron Equation

or,

New Section 1 Page 55


where for a pure species at temperature T,

ΔH = latent heat = enthalpy change accompanying the phase change

ΔV = volume change accompanying the phase change

Psat = saturation pressure, i.e., the pressure at which the phase change occurs, which is a function only of T

Rough estimates of latent heats of vaporization for pure liquids at their normal boiling points (indicated by subscript n) are given by Trouton’s rule:

where Tn is the absolute temperature of the normal boiling point. The units of ΔHn, R, and Tn are chosen so that ΔHn/RTn is dimensionless.

Standard Heat of Reaction

A standard state is defined as the state of a substance at specified pressure, composition, and physical condition as, e.g., gas, liquid, or solid.

All conditions for a standard state are fixed except temperature, which is always the temperature of the system. Standard-state properties are
therefore functions of temperature only.

Heats of reaction are based on experimental measurements. Most easily measured are heats of combustion, because of the nature of such
reactions. A simple procedure is provided by a flow calorimeter. Fuel is mixed with air at a temperature T, and the mixture flows into a combustion
chamber where reaction occurs. The combustion products enter a water-jacketed section in which they are cooled to temperature T. Because no shaft
work is produced and the calorimeter is designed to eliminate potential- and kinetic-energy changes, the overall energy balance reduces to

∆H = Q

Thus the enthalpy change caused by the combustion reaction is equal in magnitude to the heat flowing from the reaction products to the water, and
may be calculated from the temperature rise and flow rate of the water.

The enthalpy change of reaction ΔH is called the heat of reaction. If the reactants and products are in their standard states, then the heat effect is
the standard heat of reaction.

With respect to the chemical reaction, aA + bB → lL + mM, the standard heat of reaction at temperature T is defined as the enthalpy change when
a moles of A and b moles of B in their standard states at temperature T react to form l moles of L and m moles of M in their standard states
at the same temperature T.

New Section 1 Page 56


When a heat of reaction is given for a particular reaction, it applies for the stoichiometric coefficients as written. If each stoichiometric
coefficient is doubled, the heat of reaction is doubled.

For example, two versions of the ammonia synthesis reaction are written:

The symbol ΔH2980 indicates that the heat of reaction is the standard value for a temperature of 298.15 K (25°C) and for the reaction as written.

Standard Heat of Formation

A formation reaction is defined as a reaction that forms a single compound from its constituent elements so, the heat associated with this reaction
is called heat of formation and if the reactants and products are in their standard states, then the heat effect is the standard heat of formation.

For example, the formation reaction of methanol can be written as:

The usual choice for this reference temperature is 298.15 K or 25°C. The standard heat of formation of a compound at this temperature is

represented by the symbol ∆ࡴࢌ૛ૢૡ . The degree symbol denotes the standard-state value, subscript f identifies a heat of formation, and the 298 is the
rounded absolute temperature in kelvins.

For the elemental gases like H2, O2, N2 etc. the heat of formation will be equal to zero.

Standard Heat of Combustion

A combustion reaction is defined as a reaction of an element or compound with oxygen to form specified combustion products. For organic
compounds consisting only of carbon, hydrogen, and oxygen, the products are carbon dioxide and water, but the state of the water may be either
vapor or liquid.

Temperature dependence of ΔH 0

A general chemical reaction can be written as:

New Section 1 Page 57


where ʋi is a stoichiometric coefficient and Ai stands for a chemical formula. The species on the left are reactants; those on the right, products. The
sign convention for ʋi is as follows:

positive (+) for products and negative (െ) for reactants

For example, when the ammonia synthesis reaction is written:

then,

This sign convention allows the definition of a standard heat of reaction to be expressed mathematically by the simple equation:

where, ∆ࡴ૙࢏ is the enthalpy of species i in its standard state and the summation is over all products and reactants.

If the standard-state enthalpies of all elements are arbitrarily set equal to zero as the basis of calculation, then the standard-state enthalpy of each
compound is simply its heat of formation.

where the summation is over all products and reactants.

For standard reactions, products and reactants are always at the standard-state pressure of 1 bar. Standard-state enthalpies are therefore functions
of temperature only,

where subscript i identifies a particular product or reactant. Multiplying by ʋi and summing over all products and reactants gives:

or,

The standard heat-capacity change of reaction is defined similarly:

New Section 1 Page 58


or,

.On integration, we get

where, ∆H0 and ∆ࡴ૙૙ are the standard heats of reaction at temperature T and at reference temperature T0 respectively. This equation is more
conveniently expressed as:

If the temperature dependence of each reactant and product is given, then the integral will become:

Adiabatic Reactions

If no heat is added to the system from surroundings and no heat is removed from the system during a reaction, the reaction can be termed as
adiabatic reaction. This condition can be achieved by the insulation of the vessel.

When the products get heated up utilizing the heat liberated during the reaction, the temperature attained is known as the temperature of reaction. In
case of adiabatic reactions, this temperature is known as the adiabatic flame temperature. When the pure oxygen is supplied in theoretical amount
and there is no heat loss, the fuel burns at the maximum temperature known as the theoretical flame temperature.

Assuming ∆H1 be the enthalpy of cooling the reactants from temperature T1 to 298 K and ∆H2 be the enthalpy of heating the products from 298 K to
T (adiabatic flame temperature). We can write,
Q = ∆H1 + ∆H2 + ∆H0298
where,

Q = Enthalpy difference between products and reactants which will be 0 for adiabatic reactions.
ࢀ ࢀ
∆H1 = ∑ࡾࢋࢇࢉ࢚ࢇ࢔࢚࢙ ‫ ࢀ׬‬૚ ࢔࢏ ࡯ࡼ,࢏ ࢊࢀ = ∑ࡾࢋࢇࢉ࢚ࢇ࢔࢚࢙ ‫ ࢀ׬‬૚ ࢜࢏ ࡯ࡼ,࢏ ࢊࢀ

ࢀ ࢀ
∆H2 = ∑ࡼ࢘࢕ࢊ࢛ࢉ࢚࢙ ‫ࡼ࡯ ࢏࢔ ࢀ׬‬,࢏ ࢊࢀ = ∑ࡼ࢘࢕ࢊ࢛ࢉ࢚࢙ ‫ࡼ࡯ ࢏࢜ ࢀ׬‬,࢏ ࢊࢀ
૚ ૚

New Section 1 Page 59


8. Chemical Reaction Equilibrium
30 July 2021 17:46

Consider the following reaction:

The stoichiometric number for a species that does not participate in the reaction, that is, an inert species, is zero.

As the reaction progresses, the changes in the numbers of moles of species present are in direct proportion to the stoichiometric numbers. Thus for
the preceding reaction, if 0.5 mol of CH4 disappears by reaction, 0.5 mol of H2O also disappears; simultaneously 0.5 mol of CO and 1.5 mol of H2
are formed. Applied to a differential amount of reaction, this principle provides the equations:

or,

All terms being equal, they can be identified collectively by a single quantity representing an amount of reaction. Thus dε, a single variable
representing the extent to which the reaction has proceeded, is defined by the equation:

The general relation connecting the differential change dni with dε is therefore:

This new variable ε, called the reaction coordinate, characterizes the extent or degree to which a reaction has taken place.

Integrating this equation from an initial unreacted state where ε = 0 and ni = ni0 to a state reached after an arbitrary amount of reaction gives:

New Section 1 Page 60


or,

Summation over all species yields:

or,

where,

Thus the mole fractions yi of the species present are related to ε by:

Change in Gibbs Energy and Equilibrium Constant

At equilibrium, ∆Gt = 0

or,

 Both K and G depend on the Temperature only.

Effect of Temperature on Equilibrium Constant

Since,

New Section 1 Page 61


Since,

or,

On integration we get,

New Section 1 Page 62

You might also like