GKF Exam Lecture Eng
GKF Exam Lecture Eng
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.
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.
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.
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.
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)
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)
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 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)
(III.5)
(III.6)
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):
Fig. 6.2. Solubility diagram of the system with maximum and minimum solubility.
Fig. 3.2 Solubility curves of some salts in water. 1 - KNO3, 2 - Na2SO4 · 10H2O, 3 -
Na2SO4, 4 - Ba (NO3)2.
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.
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.
(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.
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:
(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
,
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.
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.
(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.
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.
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
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
From here one can obtain the following condition for chemical equilibrium in a
closed system:
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.
; ;
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.
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.
Since the equilibrium condition is the minimum of the free energy of the system
(dG = 0, dF = 0), we can write:
From here one can obtain the following condition for chemical equilibrium in a
closed system:
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.
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).
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.
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.