0% found this document useful (0 votes)
41 views49 pages

GKF Exam Lecture Eng

The document discusses phase equilibrium in heterogeneous systems, defining key concepts such as constituent substances, independent components, and the Gibbs phase rule, which relates the number of degrees of freedom to the number of components and phases. It also explains first and second-order phase transitions, highlighting the differences in enthalpy, entropy, and volume changes during these transitions. Additionally, the document provides insights into the unique phase behavior of water and sulfur, including their phase diagrams and critical points.

Uploaded by

afiqaliyev2007
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as DOCX, PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
41 views49 pages

GKF Exam Lecture Eng

The document discusses phase equilibrium in heterogeneous systems, defining key concepts such as constituent substances, independent components, and the Gibbs phase rule, which relates the number of degrees of freedom to the number of components and phases. It also explains first and second-order phase transitions, highlighting the differences in enthalpy, entropy, and volume changes during these transitions. Additionally, the document provides insights into the unique phase behavior of water and sulfur, including their phase diagrams and critical points.

Uploaded by

afiqaliyev2007
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as DOCX, PDF, TXT or read online on Scribd
You are on page 1/ 49

1.

Phase equilibrium
Consider a heterogeneous system consisting of a finite number of phases. Each
phase must be in such a quantity that one can speak about its properties from a
statistical point of view and neglect its surface properties, i.e. the behavior of
matter in the boundary layer is excluded from consideration. In such a system, in
general, both transitions of matter from one phase to another (aggregate
transformations, dissolution, distribution of the dissolved substance between two
solvents, etc.) are possible, as well as chemical reactions that lead to the
disappearance of some and the formation of other substances.
Each substance that can be isolated from the system and exist outside it, is called
the constituent substance of the system. For example, in an aqueous solution of
potassium chloride, constituents are H2O and KCl, although in solution the latter
substance exists in the form of K + and Cl- ions, but they cannot be separated
separately from the solution.
If there are no chemical reactions in the system, then the amount of each
constituent substance does not depend on the number of other substances. In the
case of the course of chemical reactions, the amounts of the constituents of the
substances entering the equilibrium system depend on each other, and the
composition of the phases of the equilibrium system can be determined knowing
the concentration of only a part of the constituents of the substances. The
constituent substances whose concentrations determine the composition of the
phases of a given equilibrium system are called independent components, or
components of the system. As components, any constituent substances can be
selected.
The properties of the system are determined not by what substances are chosen as
components, but by their number, i.e., number of components. The number of
components coincides with the number of constituents in the absence of chemical
reactions, and less when the reactions proceed. The number of components is equal
to the number of constituents of the system minus the number of equations
connecting the concentrations of these substances in the equilibrium system. In
other words, the number of components is equal to the smallest number of
components sufficient to determine the composition of any phase of the system.
The heterogeneous system is in equilibrium when:
1) the temperature of all the coexisting phases is the same (thermal equilibrium);
2) the pressure in all coexisting phases is the same (mechanical equilibrium);
3) the chemical potentials of each of the components in all phases are equal
(chemical equilibrium).
2. THE GIBBS PHASE RULE
When the external parameters (p, T) change, the equilibrium in the system is
violated; at the same time the concentrations of the components change or the old
ones disappear and new phases appear. Changes in the system occur until a new
equilibrium is established. Calculation of the number of degrees of freedom in the
system, depending on the number of components and on the change of external
parameters, is performed using the Gibbs phase rule (1876).
The Gibbs phase rule is the basic law of the theory of phase equilibria in
heterogeneous systems. This rule establishes the relationship between the number
of degrees of freedom C, the number of independent components of K, and the
number of phases $ for systems in thermodynamic equilibrium:
C = K-F + n (5)
Equation (5) serves as a mathematical expression for the phase rule: the number of
degrees of freedom of the equilibrium system is equal to the number of
independent components plus the number of external parameters n that affect the
state of the system, minus the number of phases in the system.
A more specific form of the phase rule equation depends on the number n, i.e.
number of external parameters that determine the state of the system. If
temperature and pressure are such parameters, then n = 2 and equation (5) takes the
form
C = K-Ф + 2 (6)
For condensed systems, the pressure can be assumed constant and does not affect
the state of the system, then n = 1 and equation (5) has the form
C = K-Ф + 1 (7)
If, in addition to temperature and pressure, the state of the system is determined by
other external parameters, then the value of n can be greater than two, being equal
to the total number of such parameters.
The Gibbs phase rule is the fundamental basis for the construction of state
diagrams and is used in their construction and work with them. When studying
complex heterogeneous systems that are in different conditions, the phase rule
allows one to unambiguously establish whether the system is in equilibrium, and if
not, what is the degree of deviation from the equilibrium state and what changes
should be expected in the system as it approaches the equilibrium state. For
example, at a certain temperature and p = const in a three-component condensed
system under equilibrium conditions, the maximum number of phases at C = 0 (the
number of degrees of freedom can not be negative) is four:
F = K - C + 1 = 3 - 0 + 1 = 4 (8)
If, under these conditions, more phases are present in the system, this indicates that
the system is not in an equilibrium state and, as it approaches it, a part of the
phases that are nonequilibrium with time will disappear.
For a one-component system whose equilibrium of external factors is influenced
only by temperature and pressure, the Gibbs phase rule is expressed by the formula
C = 3 - Ф (9)
It follows that in a one-component system, the number of phases in equilibrium
can not be greater than three. However, this does not mean that this single-
component system can form only three phases. So, water besides the usual forms
so-called hot ice, existing at high pressures. The point is only that more than three
equilibrium phases can not co-exist at the same time. Depending on the number of
phases in equilibrium, one-component systems can be divariant (Φ = 1, C = 2),
univariant (Φ = 2, C = 1) and invariant (Φ = 3, C = 0).

3. HETEROGENEOUS SYSTEMS.
A thermodynamic system consisting of parts that are different in their properties,
delimited by interfaces, is called a heterogeneous system.
The substances entering into the thermodynamic system can be in various
aggregate states: gaseous, liquid and solid, forming one or several phases. A
system consisting of several phases is heterogeneous, and the equilibrium
established in such a system is heterogeneous or phase. Examples of heterogeneous
systems: a mixture of two crystalline substances, a saturated solution of salt in
water and salt crystals, a mixture of several liquids that are hardly soluble in each
other; water and water vapor, etc.
Phase equilibrium in a heterogeneous system is characterized by certain
conditions: the equality t0-p in all phases of the system and the equality of
pressures and chemical potentials of each component in all phases:
ТI=ТII= ... = ТФ(thermal equilibrium condition), (1)
(mechanical equilibrium condition), (2)
(the chemical equilibrium condition), (3)
The upper indices are referred to the phases, and the lower ones to the components.

Phase (Ф) - part of a heterogeneous system, bounded by the interface and


characterized in the absence of an external force field by the same chemical,
physical and thermodynamic properties at all its points. Each phase is
homogeneous, but not continuous, i.e. can consist of individual crystals.
The number of phases of the system is divided into single-phase, two-phase, three-
phase and multiphase. A system can consist of one or more components. A
component is called an individual chemical. substance, the cat. is an integral part
of the system, can be isolated from it and exist independently.
The number of components (K) is the smallest number of individual chemicals
(components) required to form all phases of the thermodynamic system and the
mathematical expression of any phase.
From the definitions of the components it follows:
1. Each component can vary and exist independently of other components.
2. Not all components of the system are taken into account when calculating the
number of components. For example, in aqueous solution of common salt there are
several kinds of particles (H2O, NaCl, H+, Cl-, Na+, OH-), but two components
(H2O and NaCl).
3. If the substances forming the system do not interact with each other, then the
number of components is equal to the number of R substances in the system. For
chemical interactions, K is less than R by the number of bonds g. The value of g is
equal to the number of independent reaction equations. For example, in the system
CaCO3 = CaO + CO2, g = 1, K = 3 - 1 = 2.
By the number of components, one-component, two-component, etc. are
distinguished. system.
The state of the system is characterized by the number of degrees of freedom
(variance).
The number of degrees of freedom (C) is the number of thermodynamic
parameters that determine the state of the system, which can be arbitrarily changed
(independently of one another) without changing the number of phases in the
system.
These parameters include external factors (temperature, pressure) and internal
(concentration of components). In terms of the number of degrees of freedom,
systems are divided into invariant (C = 0), univariate (C = 1), bivariate (C = 2), etc.
For example, at constant pressure, a saturated salt solution has one degree of
freedom. Each arbitrarily chosen t0 corresponds to a strictly defined concentration
of saturated solution.

4. FIRST AND SECOND ORDER PHASE TRANSITIONS

The phase transition at constant temperature and pressure leads to phase


equilibrium, that is, the two phases coexist simultaneously. In accordance with the
conditions of phase equilibrium () and the determination of the chemical potential
for a single-component two-phase system, we find that( )
G = G(2) - G(1) = 0. (2)
Since ΔG = pΔV - TΔ S = ΔH - TΔS, (3)
then there are two variants: V = 0, ΔH = 0, ΔS = 0, (4)
V = 0, ΔU = 0, Δ S = 0. (5)
In the first case, the phase transition is called the first-order phase transition. In this
transition, the enthalpy, entropy, and volume functions have a discontinuity. In the
second case, these state functions do not have a discontinuity, but they have a
break, because the second derivatives of the change in Gibbs energy will not be
equal to zero:
Here we are already dealing with a second-order phase transition. In this case, the
heat capacity, the isothermal compressibility T, and the thermal expansion p
have discontinuities in the dependence curves on the state parameters.
Phase transitions of the first kind include, for example, transitions: liquid-crystal,
gas-liquid; to phase transitions of the second kind - a ferromagnet - a paramagnet, a
metal - a superconductor.
When carrying out an equilibrium process, when two phases coexist in the system:
the first and the second, not only the chemical potentials (Gibbs energies) are equal
in different phases, but also
dG(1) = dG(2),(7)
since Equation (7) is satisfied for the initial and final states. Then it follows from
the fundamental expression for the Gibbs energy that
V(1)dp - S(1)dT = V(2)dp - S(2)dTand (8)

an equation of the form (8) is called the Clapeyron-Clausius equation. Since the
phase transition process usually passes at constant temperature and pressure, the
latent heat of phase transformation from phase 1 to phase 2
Q (12) = Lt = Δ H (12): {ΔHevap; ΔHmelt; }. (9)
The Clapeyron-Clausius equation describes the temperature dependence of the
phase transformation. V(12) is the volume effect of the phase transformation.

5. PHASE TRANSITIONS. ONE-COMPONENT SYSTEMS.


For a one-component system whose equilibrium of external factors is influenced
only by temperature and pressure, the Gibbs phase rule is expressed by the formula
C = 3 - Ф (1)
It follows that in a one-component system, the number of phases in equilibrium
can not simultaneously be greater than three.
Depending on the number of phases in equilibrium, one-component systems can be
divariant (Φ = 1, C = 2), univariant (Φ = 2, C = 1) and invariant (Φ = 3, C = 0).
Examples.
1. A system consisting of a chemically homogeneous liquid. There is one phase, Φ
= 1, and one component, K = 1, so that C = 2. This means that the variables T and
P can have arbitrary values.
2. A system of liquid and its saturated vapor. In such a system, there are two
phases, Ф = 2, and one component, К = 1. According to the phase rule, the number
С = 1. An arbitrary value can have only one of the variables. For example, if you
set the temperature T, the pressure will be equal to the saturated vapor pressure
3. A system consisting of three different phases of one substance, for example, ice,
water and water vapor. This system has three phases, Ф = 3, and one component, К
= 1. According to the Gibbs rule, C = 0, i.e. there are no degrees of freedom: three
different phases can coexist in equilibrium only for certain values of T and P.

This conclusion is illustrated by the (P, T) -diagram shown in Fig. 1. Curve OA


shows saturated vapor pressure as a function of T. In states corresponding to points
on this curve, water and steam can be in equilibrium. If, while maintaining the
temperature constant, increase the pressure, then the equilibrium is violated and all
the steam passes into the water. Conversely, as the pressure decreases, the liquid
evaporates and turns into steam. Thus, in the region above curve OA water is
stable, and in the area below it there is steam.
The BO curve depicts the saturated vapor pressure by ice. Above this curve is
stable ice, and under it - steam.
So, water and steam can co-exist at the values of T and P on the curve OA, and ice
and steam on the curve BO. Therefore, the point corresponding to the equilibrium
of ice, water and steam must coincide with point A, in which the two curves
intersect. This point is called a triple point. In fact, it intersects three curves: BO,
OC and OA, which represents the equilibrium curve of ice and water. These three
curves divide the plane of the diagram into three parts corresponding to the regions
of stability of ice, water, and steam.
The triple point of water corresponds to the values t = 0.00750C, P = 0.006atm.
For some substances, the pressure at the triple point is above atmospheric, so that
the straight line P = 1atm passes below the triple point. In this case, the dashed
horizontal crosses only the curve AC and the transition from the solid phase to the
vapor phase occurs, bypassing the liquid phase. At atmospheric pressure, such
substances evaporate directly from the solid phase (this phenomenon is called
sublimation). The liquid phase can exist only at a sufficiently high pressure.

6. THE DIAGRAM OF THE STATE OF WATER


Water is a unique substance in many ways. One of these special properties is the
fact that solid water (ice) is less dense than liquid water just above the freezing
point. The phase diagram for water is shown in the Figure below .

Figure 13.26
Phase diagram for water.
Notice one key difference between the general phase diagram and the phase
diagram for water. In water’s diagram, the slope of the line between the solid and
liquid states is negative rather than positive. The reason is that water is an unusual
substance in that its solid state is less dense than the liquid state. Ice floats in liquid
water. Therefore, a pressure change has the opposite effect on those two phases. If
ice is relatively near its melting point, it can be changed into liquid water by the
application of pressure. The water molecules are actually closer together in the
liquid phase than they are in the solid phase.
Refer again to water’s phase diagram (Figure above ). Notice point E, labeled the
critical point. What does that mean? At 373.99°C, particles of water in the gas
phase are moving very, very rapidly. At any temperature higher than that, the gas
phase cannot be made to liquefy, no matter how much pressure is applied to the
gas. The critical pressure (Pc ) is the pressure that must be applied to the gas at the
critical temperature in order to turn it into a liquid. For water, the critical pressure
is very high, 217.75 atm. The critical point is the intersection point of the critical
temperature and the critical pressure.

7. THE DIAGRAM OF THE STATE OF SULFUR


Sulphur exists in two crystalline forms, rhombic and monoclinic. They can be
transformed from one form into another at 95.6°C (transition temperature) and at
one atm pressure. Below 95.6°C, rhombic is the stable form while above it,
monoclinic is the stable one. At 95.6°C, both forms are in equilibrium with each
other. Each form has its own characteristic melting point. Thus, under a pressure of
one atm, melting point of rhombic Sulphur (SR) is 114°C while that of monoclinic
Sulphur (SM) is 120°C. The liquid form of Sulphur (SL) undergoes notable
changes in color and viscosity when heated and ultimately boils at 444.7°C. Thus,
Sulphur can exist in four possible phases, two solids (SR and SM), one liquid (SL)
and one vapor (Sv) phase. However, all the four phases cannot coexist at the same
time since the number of phases coexisting in one component system cannot
exceed three.The phase diagram of Sulphur system is represented in Fig. 1

Curve AO is the sublimation curve of Rhombic sulphur and gives the vapour
pressure of rhombic sulphur at different temperatures. The two phases in
equilibrium are rhombic sulphur and the vapour. The equilibrium (SR↔SV) is
monovariant. Therefore, at one temperature, there can be one vapour pressure only.
The point Ois the transition temperature (95.6°C) at which rhombic sulphur
changes into monoclinic sulphur. Ois thus a triple point at which three phases, two
solids and the vapour (SR↔SM↔Sv) coexist in [5]TEMPERATURE, °C (Not to
scale)Fig. 4. The phase diagram for the Sulphur system equilibrium. Hence, it is a
non-variant point. The curve OBis the sublimation curve of monoclinic sulphur. It
gives vapour pressure of monoclinic sulphur at different temperatures. Asthe
number of phases is 2, the system is monovariant. The Clapeyron-Clausius
equation can be used quantitatively to study the variation of vapour pressure of
solid sulphur (SR or SM) with temperature. The point B is the melting point
(120°C) of monoclinic sulphur. This is another triple point at which three phases,
viz., sulphur monoclinic, liquid and vapour (SM ↔SL↔Sv)are in equilibrium.
This is a non-variant point. The curve The curve BE is the vapour pressure curve
for liquid sulphur. The two-phase equilibrium (SL↔Sv) is monovariant. The
variation of vapour pressure of liquid sulphur with temperature canbe studied,
again, with the help of the Clapeyron-Clausius equation. The curve OCis the
transition curve which gives the effect of pressure on the transition temperature of
rhombic sulphur into monoclinic sulphur. The equilibrium involved along the
curve is SR ↔SM..Both the phases are solid. The system is monovariant. Since
transformation of rhombic into monoclinic sulphur is accompanied by increase of
volume, the increase of pressure causes a rise in the transition temperature. This
can be predicted with the help of Clapeyron equation which, when applied to
SR↔SM equilibrium, may be put as-Where ∆Htrsis the molar heat of
transition.Since density of monoclinic sulphur (1.95 g cm-3) is less than that of
rhombic sulphur (2.05 g cm-3), VBis larger than VA. The right hand side of the
above equation is, therefore, positive. Hence, dP/dT should also have a positive
sign. This means that increase of pressure raises the transition temperature. The
curve OC, therefore, slopes away from the pressure axis.The curve BC is the
fusion curve for monoclinic sulphur. This gives the effect of pressure on the
melting point of monoclinic sulphur. The two-phase equilibrium (SM↔SL)along
the curve BC is univariant. As the melting of monoclinic sulphur is accompanied
by a slight increase of volume, it follows from Clapeyron-Clausitts equation that
the melting point will rise slightly by the increase of pressure. The curve BC,
therefore, slopes slightly away fromthe pressure axis. As the slope of this curve is
much less than that of the curve OC, the twocurves meet at the point C. Thus, C is
another triple point where three phases, viz., rhombic sulphur, monoclinic sulphur
and liquid sulphur (SR↔SM↔SL)are in equilibrium and the system is non-
variant. At C, the temperature is 151°C and the pressure is 1290 atm.
The curve CD is the fusion curve for rhombic sulphur: The equilibrium along this
curveis SR↔SL. As the number of phases is 2, the-system is monovariant.
8. THE CLAUSIUS-CLAPEYRON EQUATION
Phase transformations or transitions are the transition of matter from one phase
to another without a chemical reaction. Examples of such transformations are the
processes of the transition of matter from the liquid phase to the gaseous phase
(evaporation) and vice versa (condensation):
liquid phase gaseous phase;
the transition of matter from the solid phase to the liquid phase (melting) and vice
versa (crystallization):
solid phase liquid phase;
transition of matter from the solid phase to gaseous (sublimation or sublimation)
and vice versa (desublimation):
solid phase gaseous phase;
polymorphic transformations:
solid phase (modification 1) solid phase (modification 2)
The main parameters of such processes are pressure and temperature. These
parameters relate the Clausius-Clapeyron equation.
Consider the transition of one mole of any substance from phase 1 to phase 2.
The change in the energy of phase 1 due to the reduction in matter in it is:
dG1 = V1dP – S1dT (3.1)
The change in Gibbs energy of phase 2 due to the addition of matter to it will
be:
dG2 = V2dP – S2dT (3.2)

The total change in the Gibbs energy as a result of the phase transition is:
dG = dG2– dG1 (3.3)
If the matter is transferred from one phase to another phase in an equilibrium
way, then for the equilibrium conditions the total change in the Gibbs energy is
zero dG = 0 or dG1 = dG2. With this in mind, you can write:
V2dP – S2dT= V1dP – S1dT (3.4)
or
dP S2 −S1
=
dT V 2 −V 1 (3.5)
Since phase transitions occur at constant pressure and temperature,
ΔH
S2 – S1 = ΔS = T (3.6)

Taking equation (3.6) into account, equation (3.5) takes the form:
dP ΔH 1
= ⋅
dT T V 2 −V 1 , (3.7)
Where ΔH is the thermal effect of the phase transformation, J / mol;
V2 is the volume of one mole of the substance in phase 2;
V1 is the volume of one mole of the substance in phase 1.
Equation (3.7) was called the Clausius-Clapeyron equation. It connects the
main parameters of phase transformations-pressure and temperature. This equation
is applicable to any phase transformations.

9. APPLICATION OF THE CLAPEYRON-CLAUSIUS EQUATION TO


DETERMINE THE HEAT OF A PHASE TRANSITION.

The vapor is called saturated if it is in equilibrium with the liquid.


For the evaporation process, the Clausius-Clapeyron equation can be written in
the form:
dP ΔH исп. 1
= ⋅
dT T V пар−V жид. (3.8)

where P is the saturated vapor pressure of the liquid;


ΔНисп - heat of evaporation of a liquid;
Vpar - the volume of one mole of substance in the gas phase;
Vzhid- the volume of one mole of substance in the liquid phase.

If we consider pairs with values P and T sufficiently far from the critical point,
the volume of one mole of the substance in the gaseous form far exceeds the
volume of one mole of the substance in the liquid state, i.e. V par>>Vee. Therefore,
with a sufficient degree of approximation, we can assume
Vpar- Vision ¿ Vpar. (3.9)

Then equation (3.8) takes the form

d ln P ΔH исп 1
= ⋅
dT Т V пар (3.10)
Considering saturated vapor as an ideal gas, for one mole steam can be written:
RT
Vпар = P (3.11)
In view of equation (3.11), equation (3.10) can be reduced to the form:
d ln P ΔH исп dT
= ⋅¿ ¿ 2
dT R T (3.12)
Equation (3.12) expresses the dependence of the saturated vapor pressure on the
temperature. After separation of variables and indefinite integration, we get;
ΔН исп dT
∫ R
⋅ 2 +C
T
lnP = , (3.13)
where C is the integration constant.
In a narrow temperature range, the thermal effect of the evaporation process can
be considered independent of temperature (ΔHis = Const). Then after integration
we obtain:
ΔН 1
⋅ +C
lnP = - R T (3.14)

A definite integration of equation (3.13) in a narrow temperature interval T 1 - T2


gives:

P1 ΔH исп T 2 −T 1
ln = ⋅ ¿
P2 R T !⋅¿ T
2, (3.15)
where P1 is the saturated vapor pressure of the liquid at temperature T1;
P2 is the saturated vapor pressure of the liquid at temperature T2.
Equation (3.14) shows that in a narrow temperature range there is a linear
1
dependence of lnP on Т (Fig. 3.1):
lnP

βα

1
Т

Figure 3.1 - Dependence of lnP on saturated steamliquid in a narrow temperature


range
1
By the tangent of the slope of the line to the axis Т , one can determine the
thermal effect of the evaporation process of the liquid:
ΔН
=−tg α=tg β
- R and ΔH =Rtg α (3.16)
In a wide range of temperatures, it is necessary to take into account the
dependence of the thermal effect of evaporation on temperature.

10. HEAT OF EVAPORATION, TROUTON'S RULE.


Trouton's rule says that for many (but not all) liquids, the entropy of vaporization
is approximately the same at ~85 J mol −1K−1. The (partial) success of the rule is due
to the fact that the entropy of a gas is considerably larger than that of any liquid.
Sgas≫Sliquid(1)
Therefore, the entropy of the initial state (e.g. the liquid) is negligible in
determining the entropy of vaporization
ΔSvap=Sgas−Sliquid≈Sgas(2)
When a liquid vaporizes its entropy goes from a modest value to a significantly
larger one. This is related to the ratio of the enthalpy of vaporization and the
temperature of the transition:
ΔSvap=ΔHvap/T(3)
ΔSvap is found to be approximately constant at the boiling point (Figure 1):
ΔSvap≈85Jmol−1K−1(4)
This is Trouton’s rule, which is valid for many liquids (e.g, the entropy of
vaporization of toluene is 87.30 J K−1 mol−1, that of benzene is 89.45 J K −1 mol−1,
and that of chloroform is 87.92 J K−1 mol−1). Because of its convenience, the rule is
used to estimate the enthalpy of vaporization of liquids whose boiling points are
known.

Figure 1: Enthalpies of melting and boiling for pure elements versus temperatures
of transition, demonstrating Trouton's rule (blue data). (CC BY-SA 3.0; Mgibby5).
Trouton’s rule can be used to estimate the enthalpy of vaporization of liquids
whose boiling points are known.
Experimental values vary rather more than this and for gases such as neon,
nitrogen, oxygen and methane whose liquids all boil below 150 K, have values that
are in the range 65−75, benzene, many 'normal' liquids and liquid sodium, lithium
and iodine, in the range 80−90 and ethanol, water, hydrogen fluoride in the range
105−115 J mol−1 K−1. Thus is nothing unusual about 150 K, but rather an influence
from intermolecular interactions.
The value of ~85 J mol−1K−1 corresponds to a interaction energy of ~9.5kT per
molecule and so the boiling point gives an indication of the strength of the
cohesive energy holding molecules together in the condensed phase. When the
cohesive energy exceeds this value, as in water, then the ratio ΔHvap/T
(Equation 3) is larger and conversely the ratio is smaller when the cohesive energy
is less as in Neon or methane. The ≈9.5kT minimum energy per molecule is quite a
modest energy; if a molecule has six near neighbors this corresponds to about
3kT/2 per interaction between two molecules, roughly the average thermal energy.
Melting
There is no universal rule for the entropy of melting since a similar approximate
like that used for Trouton's law (Equation 2) does not exist. However, if the mature
of the interactions are consistent between a set of solids, then a crude correlation
can be identified (Figure 1; orange symbols).
Trouton Rule's does not Apply to Structured Liquids. For example, the entropies of
vaporization of water, ethanol, formic acid and hydrogen fluoride are far from the
predicted values. However, if the liquid presents hydrogen bonding or any other
kind of high ordered structure, its entropy will be particularly low and the entropy
gain during vaporization will greater, too. The enthalpy of vaporization is greater
for hydrogen-bonding molecules than for plain alkanes. For low-molecular weight
alcohols, this effect is pronounced. The longer the alkane chain becomes, the more
the compound behaves like a pure alkane.
Keeping in mind the relative molecular weights of the compounds, you can see
there is a decreasing effect of the hydrogen bonding (and other) effects on the n-
alcohol series as we move to larger chains and become less alcohol-like (structured
liquid) and more alkane-like (unstructure liquid). This is much more obvious when
ΔH is normalized to a per-carbon basis.
Trouton's rule hardly works for high ordered substances exhibiting hydrogen
bonding. Other factors like the enthalpy of vaporization for a long chained organic
molecule {strength of Van der Waals forces} may also play some significance role.
11. SOLUTIONS. METHODS OF EXPRESSING THE CONCENTRATION
OF SOLUTIONS.
The existence of absolutely pure substances is impossible - every substance
necessarily contains impurities, or, in other words, every homogeneous system is
multicomponent. If the impurities present in the substance within the accuracy of
the description of the system do not affect the properties under study, the system
can be considered one-component; otherwise the homogeneous system is
considered a solution.
A solution is a homogeneous system consisting of two or more components, the
composition of which can continuously vary within certain limits without a sudden
change in its properties.
The solution can have any aggregate state; respectively, they are divided into solid,
liquid and gaseous (the latter are usually called gas mixtures). Typically, the
components of the solution are separated into a solvent and a solute. Typically, the
solvent is considered to be a component present in the solution in a predominant
amount or a component that crystallizes first when the solution is cooled; if one of
the components of the solution is a liquid in a pure form, and the rest are solids or
gases, then the liquid is considered a solvent. From the thermodynamic point of
view, this division of the components of the solution does not make sense and is
therefore conditional.
One of the most important characteristics of a solution is its composition,
described by the concept of solution concentration. Below is a definition of the
most common ways of expressing the concentration and the formula for converting
one concentration to another, where the subscripts A and B refer to the solvent and
the dissolved substance respectively.
The molar concentration of C is the number of moles of νB dissolved in one liter of
the solution.
Normal concentration N is the number of moles of equivalents of a solute (equal to
the number of moles νB multiplied by the equivalence factor f) in one liter of the
solution.
The molar concentration m is the number of moles of solute in one kilogram of the
solvent.
Percentage concentration ω is the number of grams of solute in 100 grams of
solution.

(III.1)

(III.2)
(III.3)
Another way to express the concentration is the mole fraction X - the ratio of the
number of moles of this component to the total number of moles of all components
in the system

(III.4)

12. FORMATION OF SOLUTIONS. SOLUBILITY


The concentration of the component in the solution can vary from zero to some
maximum value, called the solubility of the component. Solubility S is the
concentration of the component in a saturated solution. A saturated solution is a
solution in equilibrium with the dissolved substance. The magnitude of solubility
characterizes the equilibrium between the two phases, therefore all factors that bias
this equilibrium (in accordance with the Le Chatelier-Brown principle) affect it.
The formation of a solution is a complex physicochemical process. The process of
dissolution is always accompanied by an increase in the entropy of the system;
when forming solutions, there is often a selection or absorption of heat. The theory
of solutions should explain all these phenomena. Historically, two approaches to
the formation of solutions - a physical theory, the foundations of which were laid
in the XIX century, and the chemical, one of the founders of which was DI
Mendeleev. The physical theory of solutions considers the dissolution process as
the distribution of particles of a dissolved substance between the solvent particles,
suggesting the absence of any interaction between them. The only driving force of
such a process is an increase in the entropy of the system ΔS; there are no thermal
or volume effects during dissolution (ΔH = 0, ΔV = 0, such solutions are usually
called ideal). The chemical theory considers the process of dissolution as the
formation of a mixture of unstable chemical compounds of variable composition,
accompanied by a thermal effect and a change in the volume of the system
(contraction), which often leads to a sharp change in the properties of the dissolved
substance (for example, dissolution of colorless copper sulphate CuS0 4 in water
leads to the formation of a colored solution, from which not CuS0 4 is isolated, but
the blue crystalline hydrate CuS04 · 5H2O). Modern thermodynamics of solutions
is based on the synthesis of these two approaches.
In the general case, during dissolution, the properties of both the solvent and the
dissolved substance change, which is due to the interaction of the particles with
each other for various types of interaction: van der Waals (in all cases), ion dipole
(in solutions of electrolytes in polar solvents), specific interactions (formation of
hydrogen or donor-acceptor bonds). Accounting for all these interactions is a very
difficult task. Obviously, the greater the concentration of the solution, the more
intense the interaction of the particles, the more complex the structure of the
solution. Therefore, the quantitative theory was developed only for ideal solutions,
which include gas solutions and solutions of nonpolar liquids, in which the
interaction energy of dissimilar EA-B particles is close to the interaction energies
of the identical particles EA-A and EB-B. Ideally, we can also consider infinitely
dilute solutions in which the interaction of the particles of the solvent and the
dissolved substance with one another can be neglected. The properties of such
solutions depend only on the concentration of the dissolved substance, but do not
depend on its nature.

13. SOLUBILITY OF GASES IN GASES


The gaseous state of matter is characterized by a weak interaction between the
particles and large distances between them. Therefore, gases mix in any ratio (at
very high pressures, when the density of gases approaches the density of liquids,
limited solubility can be observed). Gas mixtures are described by Dalton's law:
The total pressure of the gas mixture is equal to the sum of the partial pressures of
all the gases entering it.

(III.5)

(III.6)

14. SOLUBILITY OF GASES IN LIQUIDS


The solubility of gases in liquids depends on a number of factors: the nature of the
gas and liquid, pressure, temperature, the concentration of substances dissolved in
the liquid (the concentration of electrolytes is particularly affected by the solubility
of gases).
The nature of substances exerts the greatest influence on the solubility of gases in
liquids. Thus, in 1 liter of water at t = 18°C and P = 1 atm. is dissolved 0.017 liter.
nitrogen, 748.8 l. ammonia or 427.8 liters. hydrogen chloride. Anomalously high
solubility of gases in liquids is usually due to their specific interaction with the
solvent-the formation of a chemical compound (for ammonia) or dissociation in
solution into ions (for hydrogen chloride). Gases, whose molecules are nonpolar,
dissolve, as a rule, better in nonpolar liquids - and vice versa. The dependence of
the solubility of gases on pressure is expressed by the Henry-Dalton law:
The solubility of a gas in a liquid is directly proportional to its pressure over the
liquid.
Here C is the concentration of the gas solution in the liquid, k is the proportionality
coefficient, which depends on the nature of the gas. The Henry-Dalton law is valid
only for dilute solutions at low pressures, when gases can be considered ideal.
Gases that are capable of specific interaction with a solvent are not subject to this
law.
The solubility of gases in liquids depends strongly on temperature; the given
dependence is determined quantitatively by the Clapeyron-Clausius equation (here
X is the molar fraction of the gas in the solution, λ is the thermal effect of
dissolving 1 mole of gas in its saturated solution):

As a rule, when the gas is dissolved in a liquid, heat is released (λ <0), so with
increasing temperature the solubility decreases. The solubility of gases in a liquid
strongly depends on the concentration of other solutes. The dependence of the
solubility of gases on the concentration of electrolytes in a liquid is expressed by
the Sechenov formula (X and Xo is the solubility of gas in a pure solvent and
electrolyte solution with a concentration of C):

15. MUTUAL SOLUBILITY OF LIQUIDS


Depending on the nature of the liquid can be mixed in any ratio (in this case, talk
about unrestricted mutual solubility), be practically insoluble in each other (water -
kerosene, water - mercury, etc.); or have limited solubility (for example, water -
aniline, water - phenol, methyl alcohol - normal hexane). Let us consider the latter
case using the example of the aniline-water system. If approximately equal
amounts of water and aniline are mixed, the system will consist of two layers of
liquid; the top layer is an aniline solution in water, the bottom layer is a solution of
water in aniline. For each temperature, both solutions have a well-defined
equilibrium composition, independent of the amount of each of the components.
The dependence of the concentration of solutions on temperature is usually
represented graphically by means of a diagram of mutual solubility. This diagram
for the aniline-water system is shown in Fig. 6.1. The area under the curve is the
region of delamination of liquids.
An increase in temperature leads to an increase in the concentration of each of the
solutions (an increase in mutual solubility), and at a certain temperature, called the
critical temperature of the stratification (Tcr in Fig. 6.1), the mutual solubility of
water and aniline becomes unlimited. The aniline-water system refers to so-called
systems with an upper critical temperature of stratification; There are also systems
for which an increase in temperature leads to a decrease in the mutual solubility of
the components.

Fig. 6.1. Solubility diagram of aniline-water system


In some systems, for example in the water-diethylamine system, mutual solubility
increases with decreasing temperature, and a state of complete mutual solubility
can be achieved. The temperature below which the components are mixed together
in any relative amounts is called the lower critical dissolution temperature.
There are also such systems (for example, water - isoamyl alcohol), in which the
solubility of one of the components in the other increases with an increase in
temperature in a certain interval, and the other, on the contrary, decreases. Finally,
there are systems that have both upper and lower critical dissolution temperatures,
such as the water-nicotine system (Figure 6.2). However, the last three types of
systems are relatively rare. Usually, increasing the temperature increases the
mutual solubility of liquids.

Fig. 6.2. Solubility diagram of the system with maximum and minimum solubility.

16. SOLUBILITY OF SOLIDS IN LIQUIDS


The solubility of solids in liquids is determined by the nature of the substances
and, as a rule, substantially depends on the temperature; The information on
solubility of solid bodies is entirely based on experimental data. A qualitative
generalization of the experimental data on solubility is the "similar in like"
principle: polar solvents dissolve polar substances well and poorly - nonpolar ones,
and vice versa.

Fig. 3.2 Solubility curves of some salts in water. 1 - KNO3, 2 - Na2SO4 · 10H2O, 3 -
Na2SO4, 4 - Ba (NO3)2.

The dependence of the solubility of S on temperature is usually depicted


graphically in the form of solubility curves (Figure 3.2). Since the heat of
dissolution of solids in liquids can be either positive or negative, solubility with
increasing temperature can increase or decrease (according to the Le Chatelier-
Brown principle).

17. IDEAL SOLUTIONS


If the components of the binary solution (consisting of two components) are
volatile, the vapor above the solution will contain both components (the relative
content of the components in the vapor will usually differ from their content in the
solution-the vapor is relatively richer in the component whose boiling point is
lower). Consider a binary solution consisting of components A and B that are
unrestrictedly soluble in each other. The total vapor pressure, according to the first
Raoult law, is
(III.12)
Thus, for ideal binary solutions, the dependence of the total and partial pressure of
saturated vapor on the composition of the solution, expressed in mole fractions of
component B, is linear at any concentration (Fig. 3.3). Such systems include, for
example, benzene-toluene, hexane-heptane systems, mixtures of isomeric
hydrocarbons, etc.
Fig. 3.3. The dependence of the partial and total vapor pressures of an ideal
solution on the concentration

18. REAL SOLUTIONS


For real solutions, these dependences are curvilinear. If the molecules of a given
component interact with each other more strongly than with molecules of another
component, then the true partial vapor pressures over the mixture will be greater
than those calculated by the first Raoult law (positive deviations). If the
homogeneous particles interact with each other weaker than the heterogeneous
ones, the partial vapor pressures of the components will be less than the calculated
ones (negative deviations). Real solutions with positive deflections of vapor
pressure are formed from pure components with heat absorption (ΔP rains> 0),
solutions with negative deviations are formed with the release of heat (ΔP roducts<0).

Fig. The dependence of the partial and total vapor pressures of ideal (dashed line)
and real (solid line) binary solutions on composition with positive (left) and
negative (right) deviations from Raoult's law.

19. PRESSURE OF SATURATED STEAM OF DILUTE SOLUTIONS


If you look review the concepts of colligative properties, you will find that adding
a solute lowers vapor pressure because the additional solute particles will fill the
gaps between the solvent particles and take up space. This means less of the
solvent will be on the surface and less will be able to break free to enter the gas
phase, resulting in a lower vapor pressure. There are two ways of explaining why
Raoult's Law works - a simple visual way, and a more sophisticated way based on
entropy. Below is the simple approach.
Remember that saturated vapor pressure is what you get when a liquid is in a
sealed container. An equilibrium is set up where the number of particles breaking
away from the surface is exactly the same as the number sticking on to the surface
again.

Figure 2: Dynamic equilibrium between volatile molecules in the liquid and gas
phase.
Now suppose solute molecules were added so that the solvent molecules occupied
only 50% of the surface of the solution.

Figure 1.
A certain fraction of the solvent molecules will have sufficient energy to escape
from the surface (e.g., 1 in 1000 or 1 in a million). If you reduce the number of
solvent molecules on the surface, you are going to reduce the number which can
escape in any given time. But it will not make any difference to the ability of
molecules in the vapor to stick to the surface again. If a solvent molecule in the
vapor hits a bit of surface occupied by the solute particles, it may well stick. There
are obviously attractions between solvent and solute otherwise you would not have
a solution in the first place.
The net effect of this is that when equilibrium is established, there will be fewer
solvent molecules in the vapor phase - it is less likely that they are going to break
away, but there is not any problem about them returning. However, if there are
fewer particles in the vapor at equilibrium, the saturated vapor pressure is lower.

20. RAOULT’S LAW


Let us imagine that some substance B has been introduced into the equilibrium
system of A-vapor liquid. When the solution is formed, the mole fraction of
solvent XA becomes less than unity; The equilibrium in accordance with the Le
Chatelier-Brown principle shifts towards the condensation of substance A, i.e. in
the direction of reducing the pressure of saturated steam RA. Obviously, the
smaller the mole fraction of component A in the solution, the less the partial
pressure of its saturated vapor above the solution. For some solutions, the
following regularity, called the first law of Raoul:
The partial pressure of the saturated vapor of the solution component is directly
proportional to its molar fraction in the solution, the proportionality coefficient
being equal to the saturated vapor pressure over the pure component.
(III.10)
Since the sum of the mole fractions of all the components of the solution is unity,
for a binary solution consisting of components A and B, the following relation is
easily obtained, also the formulation of the first Raoult law:

(III.11)
The relative decrease in the vapor pressure of the solvent over the solution is equal
to the molar fraction of the solute and does not depend on the nature of the
dissolved substance.
Solutions for which Raoult's law is satisfied are called ideal solutions. Ideal at any
concentration are solutions whose components are close in physical and chemical
properties (optical isomers, homologs, etc.) and the formation of which is not
accompanied by volume and thermal effects. In this case, the forces of
intermolecular interaction between homogeneous and dissimilar particles are
approximately the same, and the formation of the solution is due only to the
entropy factor. Solutions whose components differ significantly in their physical
and chemical properties are subject to Raoult's law only in the region of
infinitesimal concentrations.

21. ELECTROLYTES. ELECTROLYTE SOLUTIONS.


An electrolyte is a substance that, in a solution or melt, is completely or partially in
the form of charged particles — ions. The process of decomposition of molecules
into positively and negatively charged ions - cations and anions - is called
electrolytic dissociation.
The laws of Raoult and the Van't Hoff principle do not hold for solutions (even
infinitely diluted) that conduct an electric current - solutions of electrolytes.
Summarizing the experimental data, Van't Hoff came to the conclusion that
electrolyte solutions always behave as if they contain more particles of a solute
than follows from the analytical concentration: an increase in boiling point, a
decrease in freezing temperature, an osmotic pressure for them is always greater
than the calculated. To account for these deviations, Vant-Goff introduced the
correction-isotonic coefficient (i) into the equations for electrolyte solutions:
 iCRT ;
Т замiKm ;
T кип iEm .
To explain the characteristics of the properties of electrolyte solutions, S. Arrenius
proposed a theory of electrolytic dissociation based on the following postulates:
electrolytes in solutions are decomposed into ions — they dissociate; dissociation
is a reversible equilibrium process; the forces of interaction of ions with solvent
molecules and with each other are small (ie, solutions are ideal). Despite the
simplicity (the reasons for dissociation are not considered, the forces of interaction
between particles are not taken into account, the formation of solvates is not taken
into account, etc.), the theory of electrolytic dissociation has made it possible to
explain a number of experimental facts.
To assess the completeness of dissociation in the theory of electrolytic
dissociation, we introduce the notion of the degree of dissociation α, which is equal
to the ratio of the number of molecules n that decay into ions to the total number of
molecules N:

The degree of dissociation depends on the nature of the solvent and the solute, the
concentration of the solution and the temperature. In terms of the degree of
dissociation, electrolytes are divided into three groups: strong (α ≥ 0.3), medium
strength (0.03 <α <0.3), and weak (α ≤ 0.03).
The process of dissociation of weak electrolytes is reversible and in the system
there is a dynamic equilibrium, which can be described by an equilibrium constant
expressed through the concentrations of formed ions and non-dissociated
molecules, called the dissociation constant (K). For some electrolyte, disintegrating
in solution into ions in accordance with the equation:
AaBb ↔ aAx- + bBy +
The dissociation constant can be expressed by the following relationship:
For the binary (decaying into two ions) electrolyte, the expression can be written in
the form:

Since the concentration of each ion for a binary electrolyte is equal to the product
of the dissociation degree α by the total concentration of the electrolyte C (αC), we
obtain the mathematical expression of the Ostwald dilution law:

For dilute solutions, α → 0, therefore (1 - α) → 1. Then we get:

The degree of dissociation of a weak electrolyte is related to the isotonic


coefficient by the following relationship:

For a binary electrolyte (n = 2), we get:  i  1 . This ratio allows, experimentally


determining the isotonic ratio of the solution, to calculate the degree of
dissociation of a weak electrolyte.
Arrhenius' assumption that in the solution of a strong electrolyte there is also a
dynamic equilibrium between molecules and ions, like in weak electrolytes, proved
to be erroneous. Experimental studies have shown that, firstly, the magnitude of
the dissociation constant of a strong electrolyte depends on the concentration (ie,
the law of effective masses is not applicable to solutions of strong electrolytes)
and, secondly, no methods were able to detect non-dissociated molecules in
solutions of strong electrolytes. This allowed us to conclude that strong electrolytes
in solutions of any concentration completely dissociate into ions and, therefore, the
patterns obtained for weak electrolytes cannot be applied to strong electrolytes
without appropriate corrections.

22. ELECTRICAL CONDUCTIVITY OF ELECTROLYTE SOLUTIONS


Depending on the nature of the conductor, the carriers of electric current, i.e.
particles that carry the charge in the system can be electrons or ions. In the first
case, when the electrical conductivity is provided by electrons, the substances are
conductors of the first kind, metals refer to them. Conductors of the second kind
include solutions of electrolytes, melts, solid electrolytes, ionized gases. The
transfer of charges in such systems is associated with the transfer of matter, so the
passage of an electric current can cause changes in conductors with ionic
conductivity.
The ability of a conductor to conduct an electric current is characterized by its
electrical conductivity L, which can be defined as the amount of electricity carried
across the conductor section per unit time at a unit voltage. Electrical conductivity
is numerically inverse to conductor resistance R. Resistance

(17.1)
where ρ is the resistivity of the conductor, l is the length of the conductor, S is the
cross-sectional area of the conductor. Consequently, the electrical conductivity

, (17.2)
where k = 1/ρ is the electrical conductivity.
In the case of solutions of electrolytes, S is the area of the electrodes between
which the solution is located, and l is the distance between them.
The specific electric conductivity ρ is equal to the electrical conductivity of a
conductor with unit sizes (S = 1, l = 1), it is expressed in Om -1m-1 (basic unit). For
electrolyte solutions, the expression k is also often used in Om-1cm-1, i.e. k
represents the electrical conductivity of 1 cm3 of a solution placed between flat
parallel electrodes located 1 cm apart.
The molar electrical conductivity is the conductivity of a solution of this volume in
which one mole of electrolyte is contained and the solution is placed between flat
parallel electrodes spaced a unit distance apart.
The basic unit of molar electrical conductivity is Ohm–1 ּ m2 ּ mol– , it corresponds
to the concentration of mole/m3. In practice, Om–1 cm2 mole–1 is also often used
as a unit of molar electrical conductivity. If to express k in Om–1cm–1, and
concentration in mole/l, then

23. DEPENDENCE OF THE EQUIVALENT (MOLAR) ELECTRICAL


CONDUCTIVITY ON CONCENTRATION. KOHLRAUSCH'S LAW.
The basic unit of molar electrical conductivity is Ohm–1 ּ m2 ּ mol– , it corresponds
to the concentration of mole/m3. In practice, Om–1 cm2 mole–1 is also often used
as a unit of molar electrical conductivity. If to express k in Om–1cm–1, and
concentration in mole/l, then
When determining the numerical values of , it is important to know the molar
concentration of which particle is considered. Often indicate the electrical
conductivity of a particle, which corresponds to its chemical equivalent, i.e.
concentration is expressed in terms of normality. Therefore, the concept of the
equivalent electric conductivity * is also used:

,
where ν+ and ν– are the number of cations and anions formed during dissociation
of the molecule, and z+ and z– are their charges.
For example, the concentration of sulfuric acid can be expressed in mol eq/l, i.e.
through the concentration of the particles 1/2 H2SO4, in this case
 (H2SO4) = 2*(1/2 H2SO4).
In the case of 1–1-valent electrolytes,  and * coincide.

Fig. 2. Dependence of the molar electrical conductivity on electrolyte


concentration

The molar electrical conductivity decreases with an increase in the electrolyte


concentration (Fig. 2). In solutions of strong electrolytes, this is due to the fact that,
as the concentration increases, the inter-ion interaction increases and the velocity
of the ions decreases. In this case, the number of ions in dilute solutions does not
change, since such a volume of solution, which contains 1 mole of electrolyte, is
always taken to determine the molar electrical conductivity. At significant
concentrations, ionic association is possible, leading to a decrease in the number of
conductive particles and, consequently, to a decrease in electrical conductivity.
In the case of weak electrolytes, the degree of dissociation decreases with
increasing concentration, and the number of ions in the solution decreases.
Kohlraus experimentally established that in the region of dilute solutions, the
molar electrical conductivity decreases linearly with an increase in the square root
of the concentration (Fig. 3), which is expressed by the empirical Kohlraush
equation:
,
where o is the limiting molar electrical conductivity, i.e. electrical conductivity at
infinite dilution (  o при c  0), A is an empirical constant depending on the
nature of the solution.

24. ELECTROMOTIVE FORCE OF GALVANIC CELL


When electric current passes through the solution, electrochemical reactions occur
on the electrode surface, which are accompanied by the entrance to the electrode or
the escape of electrons from it. In inverse processes, the electrochemical reactions
occurring at the interfaces of the conductors of the first and second kind led to the
generation of an electric current.
To carry out the electrochemical process, additional conditions are necessary: the
electrons must detach from one particle and go one way to the other in a general
way. This can be achieved by replacing the direct contact between the participants
in the reaction by contacting them with two metals connected together by a
metallic conductor. In order for the electron flow to be continuous, it is also
necessary to ensure the passage of an electric current through the reaction space,
which is usually carried out by the participants themselves in the electrochemical
reaction (if they are in the ionized state) or by special compounds with high ionic
conductivity.
The device for obtaining electrical energy through electrochemical reactions is
called an electrochemical (or galvanic) element. The simplest electrochemical
element consists of two metal electrodes (conductors of the first kind), lowered
into the electrolyte solution (conductor of the second kind).
If during the passage of an electric current in different directions on the electrode
surface the same reaction proceeds, but in opposite directions, then such
electrodes, as well as electrochemical elements composed of them, are called
reversible. An example of an invertible element is the Daniel-Jacobi element (Fig.
1)
(–) Zn | ZnSO4, р-р || CuSO4, р-р | Cu (+)
During the operation of such an element, electrochemical reactions occur on the
electrodes:
Zn  Zn2+ + 2e
Cu2+ + 2e  Cu
The total equation of the reaction in the element can be represented in the form
Zn + Cu2+  Zn2+ + Cu
When an infinitely small force passes through the current element from an external
source, these reactions proceed in the opposite direction.
An example of an irreversible element is the Volta element
(-) Zn | H2SO4 | Cu (+)
During the operation of such an element, reactions occur on the electrodes:
Zn  Zn2+ + 2e
2H+ + 2e  H2,
and the reaction in the element is represented by the equation
Zn + 2H+  Zn2+ + H2
When current flows from an external source to the electrodes, other reactions
occur:
2H+ + 2e  H2
Cu  Cu2+ + 2e,
those. in the electrochemical element, copper dissolves in sulfuric acid with the
evolution of hydrogen:
Cu + 2H+  Cu2+ + H2
The most important characteristic of an electrochemical element is its
electromotive force (EMF). E is the potential difference of a correctly open
element, i.e. the potential difference between the ends of the conductors of the first
kind of the same material attached to the electrodes of the cell. In other words, the
EMF is a potential difference under equilibrium conditions when there is no
current flowing in the circuit. If the electrodes are closed, then an electric current
will flow in the circuit, and the potential difference is the voltage of the
electrochemical cell different from the EMF by the amount of voltage drop across
the internal resistance of the element.

25. DOUBLE ELECTRIC LAYER


As noted above, a double electrical layer arises at the interface between phases as a
result of charge redistribution. Knowledge of the interface structure between the
electrode and the solution is of great importance in the study of the kinetics and
mechanism of electrochemical reactions. Helmholtz proposed the first simple
model of the DES structure in 1879. DES is considered as a flat capacitor, the
plates of which arise due to the transition of ions from the metal to the solution or
in the opposite direction, depending on the chemical potential of the ions in the
two phases. Consider, for example, a copper electrode immersed in a solution of
copper sulfate. The chemical potential of copper ions in a metal at a given
temperature is constant, and in solution its value depends on the ion concentration
(activity). If at a certain concentration in the solution is greater than in the
metal, then part of the Cu2+ ions will pass into the metal on which an excess
positive charge arises. Due to electrostatic attraction in the solution, near the
interface, a negatively charged layer of sulfate ions will appear. If the chemical
potential of the copper ions in the solution is less than in the metal, the copper ions
will go into solution, forming a positively charged capacitor plate, and the metal
will be charged negatively (Fig. 3).

Fig. 3. Formation of a double electric layer

Under real conditions, thermal motion leads to the fact that some ions leave the
surface and form a diffuse part of the double layer. Thus, DES can be represented
as consisting of a dense part, or a Helmholtz layer, and a diffuse part (Fig. 4). The
thickness of the dense part d is approximately equal to the radius of the solvated
ion, and the thickness of the diffuse part, depending on concentration and
temperature, can range from several angstroms to several thousand angstroms.
Fig. 4. Distribution of ions in a double electric layer: a - ion distribution; b - a
schematic representation of excess ions; c - change in the concentration of ions
with distance from the surface; r is the distribution of the potential

The total potential о can be represented as the sum of the potentials of the dense
part d, which varies linearly with the distance, and the diffuse part of the DEL- о.
According to the Gui-Chapman theory (1910), in the diffuse region the potential
varies exponentially. In this theory, only the Coulomb interaction of counterions
with ions of the inner lining is taken into account, and the possibility of specific
adsorption of counterions under the action of non-Coulomb (van der Waals or
chemical) forces is not taken into account. This specific interaction, which is
characteristic of multiply charged ions, dye ions, surfactants, is considered in the
theory of Stern (1924). Stern introduces the notion of an adsorption potential,
which gives an additional (to Coulomb) energy of adsorption of ions.

26. TYPES OF ELECTRODES


As follows from the Nernst equation, the potential of any electrode at a given
temperature and pressure is determined by the standard potential o and the
activities of substances taking part in the electrode reaction. The standard potential
is a constant inherent in each given electrode, while the activities of the substances
involved in the reaction may be different and depend on the composition of the
reaction medium. The nature of the effect of the activity of the components of the
solution on the magnitude of the electrode potential is directly related to the nature
of the electrode reaction. As mentioned earlier, electrochemistry uses the concept
of a reversible electrode, an electrode on which an equilibrium is established
between two oppositely directed electrode reactions, the velocities of which are
quite high from a practical point of view. Depending on the device and the nature
of the electrode process, several types of reversible electrodes are distinguished.
The most common types of electrodes are:
1. Electrodes of the first kind. These include cationic electrodes, the potential of
which depends on the activity of the cations. Such electrodes establish an
equilibrium between electrically neutral particles (for example, metal atoms) and
the corresponding cations in the solution-metal (zinc, copper, silver, etc.),
amalgams (amalgams of alkali and alkaline-earth metals), a hydrogen gas
electrode. Anionic electrodes of the first kind establish an equilibrium between
electrically neutral particles and anions - chlorine, bromine, iodine electrodes.
2. Electrodes of the second kind. These electrodes consist of three phases - a
metal, it’s hardly soluble salt and a solution containing the anions of this salt. On
these electrodes, an equilibrium is established between the metal atoms and the
anions in solution as a result of two particular equilibria: between the metal and the
cation of the sparingly soluble salt and between the anion in the solid phase of this
salt and the anion in the solution.
3. Oxidation-reduction electrodes. They are made up of an indifferent metal
(usually platinum, sometimes gold or palladium) immersed in a solution that
contains an oxidized and reduced form of the same substance.
4. Ion-selective electrodes, on which membrane potentials arise.

27. Raoult's law


Let us imagine that some substance B has been introduced into the equilibrium
system of A-vapor liquid. When the solution is formed, the mole fraction of
solvent XA becomes less than unity; The equilibrium in accordance with the Le
Chatelier-Brown principle shifts towards the condensation of substance A, i.e. in
the direction of reducing the pressure of saturated steam RA. Obviously, the
smaller the mole fraction of component A in the solution, the less the partial
pressure of its saturated vapor above the solution. For some solutions, the
following regularity, called the first law of Raoul:
The partial pressure of the saturated vapor of the solution component is directly
proportional to its molar fraction in the solution, the proportionality coefficient
being equal to the saturated vapor pressure over the pure component.
(III.10)
Since the sum of the mole fractions of all the components of the solution is unity,
for a binary solution consisting of components A and B, the following relation is
easily obtained, also the formulation of the first Raoult law:

(III.11)
The relative decrease in the vapor pressure of the solvent over the solution is equal
to the molar fraction of the solute and does not depend on the nature of the
dissolved substance.
Solutions for which Raoult's law is satisfied are called ideal solutions. Ideal at any
concentration are solutions whose components are close in physical and chemical
properties (optical isomers, homologs, etc.) and the formation of which is not
accompanied by volume and thermal effects. In this case, the forces of
intermolecular interaction between homogeneous and dissimilar particles are
approximately the same, and the formation of the solution is due only to the
entropy factor. Solutions whose components differ significantly in their physical
and chemical properties are subject to Raoult's law only in the region of
infinitesimal concentrations.

28. THE CRYSTALLIZATION TEMPERATURE OF DILUTE


SOLUTIONS
The solution, unlike a pure liquid, does not solidify completely at a constant
temperature; at a certain temperature, called the crystallization start temperature,
the crystals of the solvent begin to separate and, as the crystallization proceeds, the
temperature of the solution decreases (therefore, the crystallization temperature is
always understood as the freezing point of the solution). The freezing of solutions
can be characterized by the decrease in the freezing temperature ΔTzam equal to
the difference between the freezing point of a pure solvent T ° Sd and the
temperature of the beginning of crystallization of the solution Tz:
(III.13)
Let us consider the P-T diagram of the state of the solvent and solutions of various
concentrations (Figure 3.5), in which the curve OF is the dependence of the vapor
pressure over the solid solvent, and the curves OA, BC, DE are the vapor pressure
over the pure solvent and solutions with increasing concentrations, respectively.
The solvent crystals will be in equilibrium with the solution only when the
saturated vapor pressure above the crystals and above the solution is the same.
Since the vapor pressure of the solvent above the solution is always lower than
over the pure solvent, the temperature corresponding to this condition will always
be lower than the freezing point of the pure solvent. At the same time, the decrease
in the freezing point of the solution, ΔTzam, does not depend on the nature of the
dissolved substance and is determined only by the ratio of the number of particles
of the solvent and the solute.
Fig. 3.5 Decrease in freezing point of dilute solutions
It can be shown that the decrease in the freezing point of the solution is directly
proportional to the molar concentration of the solution:
(III.14)
Equation (III.14) is called Raoul's second law. The coefficient of proportionality K
- the cryoscopic constant of the solvent - is determined by the nature of the solvent.

29. BOILING POINT OF DILUTE SOLUTIONS


The boiling point of solutions of a non-volatile substance is always higher than the
boiling point of a pure solvent at the same pressure. Consider the P - T diagram of
the state of the solvent and solutions of different concentrations (Fig. 3.5). Any
liquid-solvent or solution-boils at a temperature at which the pressure of the
saturated vapor becomes equal to the external pressure. Accordingly, the
temperatures at which the isobar P = 1 atm. crosses the curves OA, BC and DE,
representing the vapor pressure over a pure solvent and solutions with increasing
concentrations, respectively, will be the boiling points of these liquids (Figure 3.6).
The increase in the boiling point of solutions of non-volatile substances
ΔTк = Tк – T°кis proportional to the decrease in saturated vapor pressure and,
consequently, directly proportional to the molar concentration of the solution. The
coefficient of proportionality E is the ebullioscopic constant of the solvent, which
does not depend on the nature of the dissolved substance.
(III.15)
Fig. 3.6 Boiling point of dilute solutions
Thus, Raoul's second law can be formulated in the most general form as follows:
The lowering of the freezing point and the rise in the boiling point of a dilute
solution of a non-volatile substance is directly proportional to the molar
concentration of the solution and does not depend on the nature of the dissolved
substance.
The second law of Raoul is a consequence of the first; this law is valid only for
infinitely dilute solutions. The proportionality coefficients in equations (III.14 -
III.15) - ebullioscopic and cryoscopic constants - have a physical meaning
correspondingly to an increase in the boiling point and a decrease in the freezing
point of solutions with a molar concentration equal to 1 mol / kg. However, since
such solutions are not infinitely dilute, the ebullioscopic and cryoscopic constants
cannot be directly determined, and therefore belong to the so-called. extrapolation
constants.

30. OSMOTIC PRESSURE OF DILUTE SOLUTIONS


If two solutions with different concentrations are separated with a semipermeable
partition that passes through the solvent molecules, but prevents the transition of
the particles of the dissolved substance, the phenomenon of spontaneous transition
of the solvent through the membrane from the less concentrated solution to the
more concentrated one will be observed. The osmotic properties of the solution are
quantitatively characterized by the value of the osmotic pressure. The pressure that
must be applied to the solution to prevent the solvent from moving to the solution
through the membrane separating the solution and the pure solvent is the osmotic
pressure π. The osmotic pressure of ideal solutions depends linearly on the
temperature and the molar concentration of solution C and can be calculated from
equation (III.16):
(III.16)
Equation (III.16) is the so-called The Van't Hoff principle:
the osmotic pressure of the ideal solution is equal to the pressure that the dissolved
substance would exert if it were in the gaseous state at the same temperature to
occupy the same volume as the solution occupies.

31. CHEMICAL EQUILIBRIUM


As was shown above, the flow of a spontaneous process in a thermodynamic
system is accompanied by a decrease in the free energy of the system (dG <0, dF
<0). Obviously, sooner or later (we recall that there is no concept of "time" in
thermodynamics), the system will reach a minimum of free energy. The minimum
condition for some function Y = f (x) is the first derivative to be zero and the
positive sign of the second derivative: dY = 0; d 2Y> 0. Thus, the condition of
thermodynamic equilibrium in a closed system is the minimum value of the
corresponding thermodynamic potential:
Isobar-isothermal (P = const, T = const):
ΔG = 0 dG = 0, d2G> 0
Isochoric-isothermal (V = const, T = const):
ΔF = 0 dF = 0, d2F> 0
The state of a system with a minimum free energy is a state of thermodynamic
equilibrium:
Thermodynamic equilibrium is a thermodynamic state of a system that, when the
external conditions are constant, does not change with time, and this immutability
is not due to any external process.
The theory of equilibrium states is one of the sections of thermodynamics. Next,
we will consider a special case of a thermodynamic equilibrium state - chemical
equilibrium. As is known, many chemical reactions are reversible, i.e. can
simultaneously flow in both directions - direct and reverse. If we carry out a
reversible reaction in a closed system, then after a while the system will come to a
state of chemical equilibrium - the concentrations of all the reacting substances
will cease to change in time. It should be noted that the achievement by the system
of a state of equilibrium does not mean the termination of the process; The
chemical equilibrium is dynamic, i.e. corresponds to the simultaneous flow of the
process in opposite directions with the same speed. The chemical equilibrium is
mobile - any infinitesimal external action on the equilibrium system causes an
infinitesimal change in the state of the system; Upon termination of external
influence the system returns to its original state. Another important property of
chemical equilibrium is that the system can spontaneously come to a state of
equilibrium from two opposite sides. In other words, any state adjacent to the
equilibrium state is less stable, and the transition to it from the equilibrium state is
always connected with the need for external work.
The quantitative characteristic of a chemical equilibrium is the equilibrium
constant, which can be expressed through equilibrium concentrations C, partial
pressures P or mole fractions X of reacting substances. For some reaction

the corresponding equilibrium constants are expressed as follows:

The equilibrium constant is a characteristic value for every reversible chemical


reaction; the value of the equilibrium constant depends only on the nature of the
reacting substances and temperature. The expression for the equilibrium constant
for an elementary reversible reaction can be derived from kinetic representations.
Let us consider the process of establishing equilibrium in a system in which only
initial substances A and B are present at the initial instant of time. The velocity of
the direct reaction V1 is at this moment maximum, and the velocity of the inverse
V2 is zero:

As the concentration of the initial substances decreases, the concentration of the


reaction products increases; accordingly, the rate of the direct reaction decreases,
the rate of the reverse reaction increases. Obviously, after some time the rates of
direct and reverse reaction will be equalized, after which the concentrations of the
reacting substances will cease to change, i.e. a chemical equilibrium will be
established.
Assuming that V1 = V2, we can write:

Thus, the equilibrium constant is the ratio of the rate constants of the direct and
reverse reactions. Hence the physical meaning of the equilibrium constant follows:
it shows how many times the rate of the forward reaction is greater than the inverse
velocity at a given temperature and the concentrations of all reacting substances
equal to 1 mol / l.

32. RELATIONSHIPS BETWEEN EQUILIBRIUM CONSTANTS.


The equilibrium constant expressed in terms of the molar concentrations of the
substances participating in the chemical reaction for the ammonia formation
reaction can be written as follows.
In addition, the equilibrium constant can be expressed in terms of the partial
pressures of the reacting substances:
P 2NH
k Р= 3

3
РH РN 2 2

The equilibrium constant can also be expressed in terms of the mole fractions of
the reaction participants:
2
N NH
kN= 3

3
NH NN 2 2

Finally, the equilibrium constant can be expressed in terms of the number of moles
of substances participating in the reaction:
n2NH
k n= 3

3
nH nN 2 2

In the general case, kC kP kn kN, but there is a connection between them. To


establish it, it is advisable to refer to a chemical reaction recorded in a general
form:
 1 M 1   2 M 2  ... 1 N 1   2 N 2  ... .
ν ν
ν1
PN P CN CN ν2
N2
1 2

2 ν ν −μ 1 −μ 2
k P= μ
1
μ
= 1
( RT ) 1 ( RT ) 2 .. . ( RT ) ( RT ) .. .
μ1 μ2
PM1 P M2
1 2 CM C 1 M2
And then

,
where Δn is the change in the amount of a substance as a result of a chemical
reaction.
For ideal gases, their partial pressures are related to the total pressure by the
relation:
n
Pi N i P  i P
 ni ,
so:
    1   2 ... 1   2  ...
PN1 PN2 n N1 n N2  P 
kP  1 2
 1 2  
n M1 n M2   n i 
 
PM1 PM2 
1 2 1 2
or
( ) ( )
Δn Δn
P RT
k P=k n =k n
∑ ni V
Further:
ν ν ν ν
P N1 P N22 N N11 N N22 Δn
k P= μ1
1
μ2
= μ1 μ2
P
P M1 P M2 N M1 N M2

or finally:
k P k N P n

33. ISOTHERM OF A CHEMICAL REACTION


Now we consider (with some simplifications) a more rigorous thermodynamic
derivation of the expression for the equilibrium constant. To do this, it is necessary
to introduce the concept of chemical potential. It is obvious that the value of the
free energy of the system will depend both on the external conditions (T, P or V),
and on the nature and quantity of substances making up the system. In the event
that the composition of the system varies with time (ie, a chemical reaction takes
place in the system), it is necessary to take into account the effect of the
composition change on the free energy of the system. We introduce into an
arbitrary system an infinitesimal number of dni moles of the i-th component; this
will cause an infinitesimal change in the thermodynamic potential of the system.
The ratio of an infinitesimal change in the value of the free energy of the system to
an infinitesimal amount of a component introduced into the system is the chemical
potential μi of the given component in the system:

The chemical potential of a component is related to its partial pressure or


concentration by the following relationships:

Here μ° is the standard chemical potential of the component (Pi = 1 atm, Ci = 1


mol /l.). It is obvious that the change in the free energy of the system can be related
to the change in the composition of the system as follows:
Since the equilibrium condition is the minimum of the free energy of the system
(dG = 0, dF = 0), we can write:

In a closed system, the change in the number of moles of one component is


accompanied by an equivalent change in the number of moles of the remaining
components; ie, for the chemical reaction given above, the following relation
holds:

From here one can obtain the following condition for chemical equilibrium in a
closed system:

In a general form, the chemical equilibrium condition can be written as follows:

This expression is called the Gibbs-Duhem equation. Substituting in it the


dependence of the chemical potential on the concentration, we obtain:

Since Σniμi = ΔF, and Σniμ°i = ΔF °, we obtain:

For the isobaric-isothermal process, we can similarly obtain:

The expressions obtained above are the isotherms of a chemical reaction. If the
system is in a state of chemical equilibrium, then the change in the thermodynamic
potential is zero; we obtain:

Here ci and pi are the equilibrium concentrations and partial pressures of the
starting materials and reaction products.
Since for each chemical reaction the standard change in the thermodynamic
potential ΔF ° and ΔG ° is a strictly defined quantity, the product of equilibrium
partial pressures (concentrations) raised to a power equal to the stoichiometric
coefficient for a given substance in the chemical reaction equation (stoichiometric
coefficients for the initial substances are generally considered negative) is a certain
constant, called the equilibrium constant. Equations

show the relationship between the equilibrium constant and the standard change in
free energy during the reaction. The chemical reaction isotherm equation relates
the real concentrations (pressures) of the reagents in the system, the standard
variation of the thermodynamic potential during the reaction, and the change in the
thermodynamic potential upon transition from a given state of the system to an
equilibrium one. The sign ΔG (ΔF) determines the possibility of spontaneous flow
of the process in the system. In this case, ΔG ° (ΔF °) is equal to the change in the
free energy of the system upon transition from the standard state (Pi = 1 atm, Ci =
1 mol / l) to the equilibrium state. The chemical reaction isotherm equation allows
us to calculate the value of ΔG (ΔF) upon transition from any state of the system to
an equilibrium state, i.e. answer the question whether the chemical reaction will
proceed spontaneously at the given concentrations of Ci (pressure Pi) of the
reagents:

If the change in the thermodynamic potential is less than zero, the process under
these conditions will proceed spontaneously.

34. INFLUENCE OF PRESSURE AND CONCENTRATION ON


EQUILIBRIUM
Let us consider several possible cases of equilibrium displacement.
1. The starting material is added to the system. In this case

; ;
By the equation of the chemical reaction isotherm (4.23 - 4.24), we obtain: ΔF <0;
ΔG <0. A spontaneous chemical process will occur in the system, directed toward
the expenditure of initial substances and the formation of reaction products (the
chemical equilibrium shifts to the right).
2. The reaction product is added to the system. In this case
; ;
According to the chemical reaction isotherm equation, ΔF> 0; ΔG> 0. The
chemical equilibrium will be shifted to the left (towards the consumption of the
reaction products and the formation of the initial substances).
3. The total pressure (for reactions in the gas phase) has been changed.
The partial pressures of all components of Pi in this case vary to an equal degree;
The direction of the equilibrium shift will be determined by the sum of the
stoichiometric coefficients Δn.
Since the partial pressure of the gas in the mixture is equal to the total pressure
multiplied by the molar fraction of the component in the mixture (Pi = PXi), the
reaction isotherm can be rewritten in the following form (here Δn = Σ (ni)prod-
Σ(ni)out):

We assume that P2> P1. In this case, if Δn> 0 (the reaction proceeds with an
increase in the number of moles of gaseous substances), then ΔG> 0; the
equilibrium shifts to the left. If the reaction proceeds with a decrease in the number
of moles of gaseous substances (Δn <0), then ΔG <0; the equilibrium shifts to the
right. In other words, an increase in the total pressure shifts the equilibrium toward
a process that proceeds with a decrease in the number of moles of gaseous
substances. A decrease in the total pressure of the gases in the mixture (P2 <P1)
will shift the equilibrium towards the reaction proceeding with an increase in the
number of moles of gaseous substances.
It should be noted that a change in concentration or pressure, shifting the
equilibrium, does not change the value of the equilibrium constant, which depends
only on the nature of the reactants and temperature.

35. EFFECT OF TEMPERATURE ON EQUILIBRIUM


An increase or decrease in temperature means the acquisition or loss of energy by
the system and, therefore, must change the magnitude of the equilibrium constant.
We write equation (I.99) in the following form:

Differentiating expression (4.28) with respect to temperature, we obtain for the


dependence of the equilibrium constant on temperature equation (4.29) - the Van't
Hoff isobar:
Arguing in a similar way, for a process under isochoric conditions, one can obtain
the Van't Hoff isochore:

The isobar and the van't Hoff isochore link the change in the chemical equilibrium
constant with the thermal reaction effect in isobaric and isochoric conditions,
respectively. Obviously, the greater the thermal effect of a chemical reaction in
absolute magnitude, the more strongly the temperature affects the value of the
equilibrium constant. If the reaction is not accompanied by a thermal effect, then
the equilibrium constant does not depend on temperature.
Exothermic reactions: ΔH° <0 (ΔU° <0). In this case, according to (4.29, 4.30),
the temperature coefficient of the logarithm of the equilibrium constant is negative.
Increasing the temperature reduces the value of the equilibrium constant, i.e. shifts
the balance to the left.
Endothermic reactions: ΔH°> 0 (ΔU°> 0). In this case, the temperature
coefficient of the logarithm of the equilibrium constant is positive; Increasing the
temperature increases the value of the equilibrium constant (shifts the equilibrium
to the right).
Graphs of temperature equilibrium dependences of temperature for exothermic and
endothermic reactions are shown in the figure below.

Fig. 1. Dependence of the equilibrium constant on temperature.

36. ISOTHERM OF A CHEMICAL REACTION


Now we consider (with some simplifications) a more rigorous thermodynamic
derivation of the expression for the equilibrium constant. To do this, it is necessary
to introduce the concept of chemical potential. It is obvious that the value of the
free energy of the system will depend both on the external conditions (T, P or V),
and on the nature and quantity of substances making up the system. In the event
that the composition of the system varies with time (ie, a chemical reaction takes
place in the system), it is necessary to take into account the effect of the
composition change on the free energy of the system. We introduce into an
arbitrary system an infinitesimal number of dni moles of the i-th component; this
will cause an infinitesimal change in the thermodynamic potential of the system.
The ratio of an infinitesimal change in the value of the free energy of the system to
an infinitesimal amount of a component introduced into the system is the chemical
potential μi of the given component in the system:

The chemical potential of a component is related to its partial pressure or


concentration by the following relationships:

Here μ° is the standard chemical potential of the component (Pi = 1 atm, Ci = 1


mol /l.). It is obvious that the change in the free energy of the system can be related
to the change in the composition of the system as follows:

Since the equilibrium condition is the minimum of the free energy of the system
(dG = 0, dF = 0), we can write:

In a closed system, the change in the number of moles of one component is


accompanied by an equivalent change in the number of moles of the remaining
components; ie, for the chemical reaction given above, the following relation
holds:

From here one can obtain the following condition for chemical equilibrium in a
closed system:

In a general form, the chemical equilibrium condition can be written as follows:

This expression is called the Gibbs-Duhem equation. Substituting in it the


dependence of the chemical potential on the concentration, we obtain:
Since Σniμi = ΔF, and Σniμ°i = ΔF °, we obtain:

For the isobaric-isothermal process, we can similarly obtain:

The expressions obtained above are the isotherms of a chemical reaction. If the
system is in a state of chemical equilibrium, then the change in the thermodynamic
potential is zero; we obtain:

Here ci and pi are the equilibrium concentrations and partial pressures of the
starting materials and reaction products.
Since for each chemical reaction the standard change in the thermodynamic
potential ΔF ° and ΔG ° is a strictly defined quantity, the product of equilibrium
partial pressures (concentrations) raised to a power equal to the stoichiometric
coefficient for a given substance in the chemical reaction equation (stoichiometric
coefficients for the initial substances are generally considered negative) is a certain
constant, called the equilibrium constant. Equations

show the relationship between the equilibrium constant and the standard change in
free energy during the reaction. The chemical reaction isotherm equation relates
the real concentrations (pressures) of the reagents in the system, the standard
variation of the thermodynamic potential during the reaction, and the change in the
thermodynamic potential upon transition from a given state of the system to an
equilibrium one. The sign ΔG (ΔF) determines the possibility of spontaneous flow
of the process in the system. In this case, ΔG ° (ΔF °) is equal to the change in the
free energy of the system upon transition from the standard state (Pi = 1 atm, Ci =
1 mol / l) to the equilibrium state. The chemical reaction isotherm equation allows
us to calculate the value of ΔG (ΔF) upon transition from any state of the system to
an equilibrium state, i.e. answer the question whether the chemical reaction will
proceed spontaneously at the given concentrations of Ci (pressure Pi) of the
reagents:

If the change in the thermodynamic potential is less than zero, the process under
these conditions will proceed spontaneously.

37. THE INFLUENCE OF EXTERNAL CONDITIONS ON THE


CHEMICAL EQUILIBRIUM. LE CHATELIER'S PRINCIPLE
With constant external conditions, the system can be in an equilibrium state for as
long as desired. If these conditions are changed (ie, any external influence is
exerted on the system), the equilibrium is violated; a spontaneous process arises in
the system, which continues until the system again reaches the equilibrium state
(already under the new conditions).
The effect of such factors as pressure, concentration and temperature, as well as
any others, on a system in a state of equilibrium, is generalized by the principle of
equilibrium displacement, also called the Le Chatelier-Brown principle:
If an external action is exerted on a system in a state of true equilibrium, then
a self-arbitrary process arises in the system that compensates for this effect.
The Le Chatelier-Brown principle is one of the consequences of the second law of
thermodynamics and is applicable to any macroscopic systems in a state of true
equilibrium

38. CALCULATION OF THE EQUILIBRIUM CONSTANT


The constant of chemical equilibrium depends on the nature of the reactants, on
temperature and is related to the change in the standard Gibbs energy AG ° of the
chemical reaction by the equation

Equation (6.14) allows the value of AG ° to determine Kp, and hence the
equilibrium concentration. Hence it follows that the more significant the decrease
in the Gibbs energy, i.e. the more the equilibrium is shifted towards the reaction
products, the greater the value of the equilibrium constant. At high negative AG °
values, the reaction products prevail in the equilibrium mixture. If AG °> 0, then
the initial substances prevail in the equilibrium mixture.
Combining equations (6.10) and (6.14) in terms of the value AG0, we obtain
Equation (6.15) makes it possible to calculate the equilibrium constant and the
degree of equilibrium transformation from the values of A // 0 and AS0.
It is clear from equations (6.14) and (6.15) that the equilibrium constant depends to
a large extent on temperature. For endothermic processes, an increase in
temperature corresponds to an increase in the equilibrium constant, for exothermic
processes - to its decrease. The equilibrium constant does not depend on pressure
(if p is not very large).

39. Dependence of the equilibrium constant on temperature


An increase or decrease in temperature means the acquisition or loss of energy by
the system and, therefore, must change the magnitude of the equilibrium constant.
We write equation (I.99) in the following form:

Differentiating expression (4.28) with respect to temperature, we obtain for the


dependence of the equilibrium constant on temperature equation (4.29) - the Van't
Hoff isobar:

Arguing in a similar way, for a process under isochoric conditions, one can obtain
the Van't Hoff isochore:

The isobar and the van't Hoff isochore link the change in the chemical equilibrium
constant with the thermal reaction effect in isobaric and isochoric conditions,
respectively. Obviously, the greater the thermal effect of a chemical reaction in
absolute magnitude, the more strongly the temperature affects the value of the
equilibrium constant. If the reaction is not accompanied by a thermal effect, then
the equilibrium constant does not depend on temperature.
Exothermic reactions: ΔH° <0 (ΔU° <0). In this case, according to (4.29, 4.30),
the temperature coefficient of the logarithm of the equilibrium constant is negative.
Increasing the temperature reduces the value of the equilibrium constant, i.e. shifts
the balance to the left.
Endothermic reactions: ΔH°> 0 (ΔU°> 0). In this case, the temperature
coefficient of the logarithm of the equilibrium constant is positive; Increasing the
temperature increases the value of the equilibrium constant (shifts the equilibrium
to the right).
Graphs of temperature equilibrium dependences of temperature for exothermic and
endothermic reactions are shown in the figure below.

Fig. 1. Dependence of the equilibrium constant on temperature.

40. EQUILIBRIUM IN HETEROGENEOUS SYSTEMS.


In deriving the law of acting masses, it was assumed that all the reagents are
gaseous substances, and the considered reaction 3H2 + N2 = 2NH3 is a
homogeneous chemical reaction. If, however, liquid or solid substances that do not
form solutions with each other participate in the reaction, then their partial
pressures at constant temperature are constant as constant pressures of saturated
vapors of substances at a constant temperature. these constants are introduced
under the sign of the equilibrium constant, which in this case will be determined
only by the pressure of the gaseous reagent.
The effect of this provision on the example of the dissociation reaction of CaCO3:
CaCO3 (TB) ↔ CaO (TB) + CO2 (D).
In this case:

but since CaO and CaCO3 are condensed substances, then at T = const, P CaO and
PCaCO3 are also constant, therefore:
Kp = PCO2,
where PCO2 is the dissociation elasticity or dissociation pressure of CaCO3.
The temperature dependence of the dissociation elasticity of a number of
compounds is shown in the figure below.
Under chemical equilibrium conditions at an appropriate temperature, the
dissociation pressure is a completely determined quantity.
When the temperature changes, the dissociation pressure also changes. The nature
of this dependence makes it possible to conclude that the compound is stable
against decomposition.
For example, for dissociation reactions of some oxides:
2Ag2O (sd) ↔ 4Ag (sd) + O2 (g), Kp = PO2,
2CuO (sd) ↔ 2Cu (sd) + O2 (g), Kp = PO2,
2Cu2O (sd) ↔ 4Cu (sd) + O2 (g), Kp = PO2,
The points of intersection of the isobar correspond to the partial pressure of oxygen
in the earth's atmosphere (0.21 * 105 Pa), with the corresponding dissociation
elasticity curves giving temperatures above which the compounds will readily
dissociate, and at lower temperatures dissociation will be suppressed by the partial
pressure of atmospheric oxygen. Analysis of the curves also leads to the
conclusion that Cu2O is the most stable, and Ag2O is the least stable. Analogous
arguments can be made to evaluate the behavior of the same compounds in an
atmosphere of pure oxygen.

You might also like