Download as pdf or txt
Download as pdf or txt
You are on page 1of 23

Thermodynamics

System: part of the universe that is being the object of study


• open: exchanges energy and matter
• closed: only exchanges energy
• isolated: does not exchange anything
Environment: rest of the universe

What separates the system from the environment are the walls, in turn they can be:
• rigid or mobile
• permeable, semipermeable or impermeable
• adiabatic (does not allow heat exchange) or diathermic

What characterizes a system are the states, which are defined by variables:
• intensive: they do not depend on the amount of matter in the system, such as P, T, ρ, ...
• extensive: they depend on the amount of matter in the system1 (they are additive), such as m, V, ...

The variables are thermodynamic properties that are related through a state function2 f(xi):
• it has a complete derivative d f =∑
( ∂∂xf )
i x j ≠i
d xi

• it does not depend on the history of the variables ∮ df =0


∂ ∂f = ∂ ∂f
• it fullfills the Schwartz condition
( )
∂ x j ∂ xi ∂ x i ∂ x j ( )
The different processes can be arranged into:
• isotherm (T=cte)
• isobaric (P=cte)
• isochoric (V=cte)
• adiabatic (∆Q=0)
• cyclic (final state = initial state)
• reversible (takes place through an infinite succession of equilibrium states; which if reversed, both the
environment and the system return to their original states, without an extra energy consumption)
• irreversible (an infinitesimal change does NOT produce a change in the direction of the transformation;
the states can be reversed, but using additional energy to undo the changes caused in the
surroundings. A global process will be irreversible if any of the involved steps are irreversible)

Zero principle: Thermal Equilibrium


Temperature is a measure of the amount of kinetic energy that the components (particles/molecules) of a
system have; and allows us to compare amounts of energy from different systems. If we have two closed
systems A and B, where TA > TB, energy will be transferred from system A to B (in the form of heat), until T A =
TB, at which point we will say that it has been reached thermal equilibrium. At this point, energy can continue
to be transferred from one system to the other, but the net change will be zero, and the temperature will not
change.

First Principle: Work and Heat


We will use the following sign criteria: if the system does anything, it will be considered as negative in sign;
meanwhile if the system receive anything it will be considered as positive. Also, work can be of different types

1 These variables are affected when subdividing the system, changing their magnitude.
2 For example: P·V=n·R·T

1
(product of a force by a displacement):
• mechanical: dW =− P dV
• electrical: dW =E dq
• surface tension: dW =γ dA
• gravitational: dW =m g dh
• …

The work is NOT a state function. For example, consider the mechanical work of a piston expansion (ideal gas,
T = const, complete expansion against external pressure):

dx Fext δW =−F dx dV =S dx F=P S δW =−P dV


We can express the expansion in one or several steps:
1 step
initial: Po, Vo 1 3
final: ¼Po = Pext, 4 Vo |W|= Po ( 4 V o−V o )= P o V o
4 4

2 steps
initial: Po, Vo 1 1
final: ½Po = Pext, 2 Vo |W 1|= Po ( 2V o −V o )= Po V o
2 2
initial: ½Po, 2Vo 1 1
final: ¼Po = Pext, 4 Vo |W 2|= P o ( 4 V o −2 V o )= Po V o |W|=P o V o
4 2

Infinite steps: reversible process


now the change is infinitesimal and the pressure if virtually Pext at any times: PV=nRT
Vf Vf nRT Vf
|W|=∫V P ext dV =∫V
o o V
dV |W|=n R T ln ( )
Vo
=P o V o ln (4 )≈1,39 P o V o

The heat is NOT a state function and it can be measured as Q=c X m Δ T , where cX is he heat capacity or
amount of heat required to increase the temperature by one degree per unit mass of a substance. We can
express it as:

• heat capacity at constant pressure c P = ( δQ


∂T ) P
δQ
c =(
∂T )
• heat capacity at constant volume V
V

The first principle arises from the conservation of energy: the internal energy of a system can be expressed as
heat or work.
dU =δ Q+δ W ΔU =Q +W
If our conditions are:

• constant volume (in absence of mechanical work): dU =δ Q ( ∂∂UT )


cV =
V
∂H
c =(
∂T )
• constant pressure (Enthalpy, H): dH =δ Q P Δ H=Δ U + P Δ V P
P

We can relate both heat capacities using Mayer's relationship:

( ∂∂ HT ) −( ∂U
c P −c V =
P
=
∂U
∂T ) ( ∂T )V
+P (
∂V

P
∂U
∂T ) ( ∂T ) P V

∂V
dU =(
P
∂U
∂T ) dT +(
V
∂U
∂V ) dV ; dV =(
T ∂T ) dT +(
∂V
∂P) P
dP dP=0
→→ T
dV =( ∂∂ VT ) dT
P

2
∂U ∂U ∂U ∂U ∂V
( ∂U
dU P =
∂T ) dT +(
V ∂V ) ( ∂V
∂T ) dT ( ∂T )
T
=(
∂T ) +(
P ∂V ) ( ∂ T ) P V T P

∂U ∂V ∂V
c −c =(
P V
∂U
∂T ) ( ∂V ) ( ∂ T )
+
V
+ P(
∂ T ) ( ∂ T ) [( ∂ V )
T

∂U
=
P
∂U
] ∂∂VT )
+P (
P V T P

In the case of an ideal gas, the internal energy only depends on the temperature 3, dU ideal gas(T=const )=0 ,
and therefore:

c P −c V =P ( ∂∂ VT ) P
PV =nRT
→ → ( ∂∂ VT ) = nPR
P
c P−c V =nR

∆U and ∆H for closed systems

( ∂∂UT ) dT +( ∂∂ UV ) dV =c dT +( ∂∂ UV ) dV
Cyclic process (state functions):
U =f (T , V )

H =f (T , P )
V T
V

( ∂∂HT ) dT +( ∂∂ HP ) dP=c dT +( ∂∂ HP ) dP
P T
P
T

T
} (Kirchoff)

Δ U =0 , Δ H =0
Phase change (T,P=const):
Δ U =Δ H −P ΔV
Isobaric process (P=const):
dH =c P dT Δ U =Δ H− PΔ V (this applies always)
Isothermal process (T=const):
Vo
dT =0 (ideal gas)
{ ΔU =0
Δ H=Δ U +Δ(P V )=0
Q=−W
{
reversible

irreversible
W =−∫ Pgas dV =n R T ln

W =− Pext ΔV
( )
Vf

Isochoric Process (V=const):


dU =cV dT Δ H =ΔU +V Δ P (this applies always)
Adiabatic process (δQ=0 for an ideal gas):
−n R T c Tf V
d U =cV dT=δW c V dT =
V
dV ∫ TV dT =−n R∫ dV
V
c V ln ( )
To
=n R ln o
Vf ( )
nR
Tf V cP
=
To Vf ( ) o cV
γ=
cV
(adiabatic coefficient)

Heat of Reaction
It is the heat involved during the complete transformation of reactants into products, under certain conditions
of pressure and temperature.
• if heat is released: exothermic (∆H<0)
• if heat is absorbed: endothermic (∆H>0)
The standard enthalpy of formation corresponds to the heat of reaction for the formation of one mole of
3 In a monatomic ideal gas, since there is no interaction between the molecules, the internal energy is kinetic energy,
1 2 3
which can be demonstrated by kinetic theory of gases that its average is ⟨ m v ⟩= k B T . From statistical
2 2
thermodynamics the only contribution to the internal energy in this type of gas is the translational partition function:
∂ lnQ trans 1 ∂ const T
3/ 2
3
Δ U =k B T 2 =k B T 2 = k B T . If there were rotation of the ideal gas molecules, the
∂T const T
3/ 2
∂T 2
contribution to the internal energy would also be a function of T ( 3 /2 k B T or k B T for linear molecules).

3
substance from the constituent elements (Hess's Law):
H m molar enthalpy
ΔR H =∑ νi H m , i { ν stoichometric coefficient (>0 products, <0 reactants)
The experimental determination is carried out using a calorimeter (measuring temperature changes).
By convention, the standard enthalpy of formation (1 bar/atm, 25 ºC) of the elements in their reference state is
zero.

[Example] Consider an ideal gas in the chamber of a piston, where the initial volume is 2 L and the initial
pressure is 8 atm. Suppose the systems expands isothermally (i.e., the piston rises) reaching a final volume of
5.5 L, opposing a constant external pressure of 1.75 atm. Calculate W, and the final pressure of the gas.
δ W =−P dV W =−1.75 ·(5.5−2) ·101.325=−620.6 J Po · V o =Pf ·V f Pf =2 ·8/ 5.5=2.91 atm
[Example] A gas contained in a piston expands extremely slowly from 25 mL to 75 mL, while the temperature
remains constant at 25 ºC. If there are 0.001 moles, calculate the work done by the system.
δ W =n R T ln( v o /V f ) W =0.001· 8.314 ·(273.15+25)· ln(25 /75)=−2.72 J
[Example] A gas changes its volume from 4 L to 6 L against an external pressure of 1.5 atm while absorbing
1000 J of heat. What is the change in the internal energy of the system?
Q=1000 J W =−1.5·(6−4 )·101.325=−303.98 J ∆ U =Q +W =696 J
[Example] A cylinder filled with 0.04 mol of an ideal gas expands reversibly from 50 mL to 375 mL at a
constant temperature of 37 °C. In doing so, it absorbs 208 J of energy in the form of heat. Calculate ∆H.
Q=208 J W =n R T ln (V o / V f )=0.04 ·8.314 ·(273.15+37) ·ln (50/ 375)=−207.8 J
∆ U =Q+W ≈0 J ∆ H =∆ U +∆ ( PV )=∆ U +( Pf V f −P o V o )=0 J
[Example] Consider 1 mole of an ideal gas at an initial pressure of 1 atm and an initial temperature of 273.15 K.
Suppose that it expands adiabatically against a pressure of 0.435 atm until its volume doubles. Calculate the
work, final temperature and ∆U of the process.
3
∆ U =Q+W =W ∆ U =−0.435 · 22.4 ·101.325=−987.3 J ∆ U =cV· ∆ T = R · ∆ T ∆ T =−79.2 deg T f =194 K
2

Heat /Work conversion


The fact that the first principle is met, it does not ensure that a process is possible or not. Some processes
occur spontaneously in one direction and not in another: a hot cup cools over time, but not the other way
around. Work can be converted entirely into heat, but the opposite is very difficult and even impossible
working cyclically (perpetual mobile).

Kelvin (1851): A process whose only result is the absorption of heat from a source and the conversion of this
heat into work is not possible.
QH W
TH engine

Clausius (1854): A process whose only result is the transfer of heat from a body of lower temperature (TLow) to
another of higher temperature (THigh) is not possible.
QL
TH TL

4
Second Principle: Entropy (Carnot Cycle)
The (thermal4) engine does a work on the surroundings.

Steam engine: hot steam moves a piston producing mechanical work. In order to close the cycle, we need to
cool it down so that it recovers its initial temperature and position. Considering an ideal gas:

Process ΔU Qrev W
V2
P A P1,V1,T1
A → B (isothermal) 0 −W −R T 1 ln
( )
V1
B P2,V2,T1 B →C (adiabatic) c V (T 2−T 1 ) 0 ΔU
V4
C → D (isothermal) 0 −W −R T 2 ln ( )
V3
C P3,V3,T2 D → A (adiabatic) c V (T 1−T 2 ) 0 ΔU
D P4,V4,T2

V V2
Total 0 R (T 1−T 2 ) ln ( )
V1
−( Q H +QC )

where the following relationship can be extracted from the adiabatic cycles:
cV cV
V2 T V4 T V 2 V1 V2 V3
( )( )
V3
= 2
T1
nR
( )( )
V1
= 1
T2
nR
→ =
V3 V4
→ =
V1 V4

Performance or efficiency is defined as the work produced in each cycle, for the energy consumed (in this case
energy for heating):
V2

η=
|W| Q H +Q L
= =
R ( T 1−T 2 ) ln( )V1
T 1 is the hot spot source η=1−
TL
QH QH V2 TH
R T 1 ln
( )
V1
which corresponds to the ideal efficiency of a Carnot engine. In addition, it can be also expressed as:
Q T QL QH Q Q
η=1+ L =1− L + =0 (including the adiabatic steps) 0= L +0+ H + 0
QH TH TL TH TL TH
which suggests the existence of another property that must be a function of state, and which we will call
entropy:
δQ
dS= rev
T

In an irreversible cyclic process, by definition, the efficiency of the heat engine will be lower:
T Q T Q Q Q Q Q Q
ηrev >ηirrev 1− C >1+ L
TH QH
− L> L
TH QH
− H> L
TH TL
0> L + H
TL TH ∑
cycle T
⩽0 → ∮irrev δTQ ⩽0

If we assume a cycle consisting on an irreversible process (A → B) followed by a reversible one (B → A):


B δQ A δQ B δQ B δQ B δQ B δQ
∮irrev δQ
T
=∫A
T
irrev
+∫B
T
rev
<0 ∫ A T
irrev
−∫ A T
rev
<0 ∫ A T
irrev
<∫ A T
rev
Δ S irrev rev
A →B < Δ S A → B

so that:

4 In the case of a refrigerator, where we transfer heat from a cold source to a hot one, work must be done on the
engine for that transfer to occur (opposite direction to the heat engine).

5
δQ
dS >spontaneous and irreversible process

dS⩾
δQ
T
{
dS=

dS <
T
δQ
T
δQ
T
reversible process

impossible process

given that in an adiabatic process the comparison is performed against zero (Q=0):
Δ Suniverse =Δ S system + Δ Ssurroundings >0
The entropy of the universe always increases.

Second Principle: Entropy (Statistical Thermodynamics)


Entropy depends on the number of microscopic combinations (or microstates) available to a molecular system
which are compatible with a given macroscopic state (often seen as the “degree of disorder”), and is
therefore related to the number of accessible combinations5 (Ω):
N!
S=k B ln Ω=k B ln
∏ ni!
where kB is the Boltzmann constant (1.380649·10-23 J/K).
The isothermal and reversible expansion to twice the initial volume of an ideal gas is characterized by:
nRT Vf
Δ U =0 δ Q rev =−δ W rev =P dV =
V
dV Q rev =n R T ln ( )
Vi
=n R T ln 2

On the other hand, if we calculate the entropy value for the process in which we initially have N particles in
half the volume, and have N/2 in each half volume, once reached the equilibrium:
N! N! N 2
Si =k B ln =k ln1=0
N !0! B
S f =k B ln
N N
! !
( )
=k B ln N!−k B ln !
2
2 2
ln N!≈ N ln N −N (Stirling approximation, huge N)
N N N N
[ 2 2 2]
S f = [ k B N ln N −k B N ]− 2 k B ln −2 k B =k B N ln ( )N/2
=n R ln 2

Since ∆S = Sf – Si, we can stablish the same relationship between entropy and the reversible heat as in the
previous section.

Second Principle: Entropy


Suppose we have two bodies in contact at different temperatures (TH, TL and TH > TL). The zero principle tells
us that a heat flow will be established from the hottest body (T H) to the coldest body (TL), until thermal
equilibrium is reached (and both share the same temperature). Then, the following relationship is
accomplished:
QH QL Q Q Q Q Q δ Q rev
< Q H =−Q ; Q L=Q − < 0< + Δ S= ; Δ S>0 dS=
TH TL TH TL TH T L T T
Thus, we define a new magnitude, which from one that is not a state function (δQ), when divided by a factor
(temperature), is transformed into a state function (S); and also allows establishing when a process is
spontaneous: ∆S>0. The heat of a reversible process between to states A and B is:
δ Q rev dU P dV ideal gas δ Q rev dU n R
δ Q rev=dU + P dV = + = + dV
T T T → T T V

5 This formula is for distinguishable particles, distributed without repetition. In the case of the "tute" card game, the
value would be 40!/(10!4) [~4.7·1021 combinations]

6
B B B B B
δ Q rev
dU nR 1 ∂U nR

A T A T
=∫
+∫
A V
dV =∫
A T ∂T
dT +∫
A V
( )
dV
B B
δQ c V T V

B
A A
( )
∫ T rev =∫ TV dT +nR ln V B =cV ln T B +nR ln V B
A A A
( ) ( )
δ Q rev
∫ T
=[ cV lnT B +n R ln V B ] −[ c V ln T A +n R lnV A ] =S (T B ,V B )−S (T A ,V A )=∆ S
A

It can be shown that the above equation holds for any system, and not just for an ideal gas:
δ f =a (x , y )dx+b (x , y )dy δ f =0 a( x , y)dx +b( x , y)dy =0 integrate g (x , y )=const

∂g
δ f =u(x , y )
∂g
( )
∂x
dx+u (x , y )
∂g
∂y
dy
a(x , y)=u( x , y)
( )
b (x , y )=u(x , y )

∂ δf≠ ∂ δf
{∂x
∂g
∂y
∂ x≠ ∂ − y
δf
u(x , y)
( )
is a state function
( )
δ f =− y dx+ x dy 1≠−1
∂x ∂ y ∂ y ∂x ∂x ∂y
dy dx y
− y dx+ x dy=0 = ln y +c 1=ln x+ c2 const =g(x , y)=
y x x
∂g ∂g y 1 δ f y 1
δ f =u ( )
∂x
dx +u
∂y ( )
dy=u − 2 dx +u
x x
dy( ) ()u=x 2
x2
=− 2 dx + dy
x x

Suppose that we can reach state B (final) from state A (initial) by doing work through a reversible process, as
well as an irreversible one. Since the internal energy of both processes must be the same (∆U = UB – UA):
dU = δ Qrev + δ W rev =δ Q irrev + δ W irrev T dS+ δ W rev =δ Q irrev + δ W irrev
δ Q irrev 1 δ Q irrev δ W lost δQ
dS=
T
+ (δ W irrev −δ W rev ) >0
T
dS=
T
+
T (
>0
)
dS≥
T
δ Qirrev δ W lost δ Q irrev
dS system =
T
+ (
T >0
)
dS surroundings =−
T
dSuniverse =dS system +dS surroundings
δ W lost
dSuniverse = ≥0
T

Entropy of common/simple reversible processes


Cyclic process:
δ Q rev
Δ S=0=∮ dS=∮
T
Adiabatic process:
δQ irrev
Q=0 Δ S=0 ∫ >0 (irreversible)
T
Isothermal process:
δQ Q
Δ S=∫ =
T T
Isobaric process:
δQ c
δQ=c P dT Δ S=∫ =∫ P dT
T T
Phase change (T,P=const):
QP Δ H
= Δ S=
T T
State change T1, V1, P1 → T2, V2, P2 (can be decomposed into an isochoric and an isobaric/isothermal
processes):

7
δQ dU c P T2 V
Δ S=∫
T
=∫
T
−∫
δW
T
=∫ V dT +∫ dV
T T
PV =nRT
→→
Δ S=c V ln ( )
T1
+nR ln 2
V1 ( )
For an irreversible process we will propose series of reversible steps that connect the initial state with the
final one:
H 2 O(l ,−10o C ) → H 2 O (s ,−10o C ) J
J c liquid
P =75,3
(1)↓ ↑(3) Δ S rev =Δ S1 +Δ S 2 +Δ S 3=−20 ,56 mol · K
(2) mol · K J
H 2 O(l ,0 o C ) H 2 O( s,0 o C)
→ c solid
P =36 ,8
mol · K
liquid
cP Δ H fusion (273)
solid
cP J
Δ S 1=∫263
273
dT Δ S 2=− Δ S 3=∫273
263
dT Δ H fusion (273)=6004
T 273 T mol
The resulting entropy for the system is negative since the water molecules are strongly arranged in the ice.
The entropy of the surroundings must be calculated, corresponding to the absorption of the heat at constant
pressure released during the solidification process (as opposed to fusion, hence the change of sign in the
previous equation):
Δ H fusion (263)
Δ S surroundings,irrev=
263
In this case it is positive, since the surroundings absorb that heat. The temperature is the one at which the
irreversible process happens: -10o C

) dT Δ Ssurroundings =21 ,35 molJ · K Δ S universe=0,79 molJ · K


263
Δ H fusion (263)=Δ H fusion (273)+∫273 ( c liquid
P −c solid
P

Maxwell cross-differentiation relations


P H S
 
H =U +PV dH =T dS +V dP
G
 
U dU =T dS−P dV
{ G=H −TS
F=U −TS
dG=−S dT +V dP
dF=−S dT−P dV
T F V

Third Principle: Spontaneity and Free Energy


We are going to try to relate the principle of spontaneity (∆S universe≥0) with some property of the system so
that we do not have to also evaluate the surroundings:
Helmholtz potential (V,T=const)
Δ F system =Δ U system−T Δ S system=Q system −T Δ S system the heat relased by the system
is absorbed by the surroundings
Δ F system =−Qsurroundings −T Δ S system =−T Δ Ssurroundings −T Δ S system
Δ F system =−T Δ S universe
so that the value of the free energy (of Helmholtz) of the system gives us information about the spontaneity
or not of a process:
Δ F system <0 spontaneous process
Δ F system =0 equilibrium
This result can also be extended to the variation of Gibbs free energy (∆G, G=F+PV)

8
Open systems: Chemical Potential
The chemical potential accompanies the variation in the number of moles of a certain substance, and is an
indication of the activity or nature of that substance.
- Gibbs-Duhem:
dU =T dS−P dV + ∑ µi dni U =T S−P V + ∑ µ i ni
dU T=T dS−P dV + ∑ µ i dni + {∑ ni d µ i−v dP }=0
RT RT
( PP )=RT ln ( a )
P
ni d µ i=V dP d µi =υ dP υ= µ i−µ oi =∫P dP µi −µio=RT ln i
P o P o

where Po corresponds to the reference pressure (1 bar/atm) and μio corresponds to the reference chemical
potential (298 K, 1 bar/atm). The P/Po quotient is replaced by the activity (ai), which allows the conditions to be
generalized (solutions, ...)
∂G
- Gibbs: it is defined as the change in free energy per mole of substance, µi =
∂ ni P,T ( )
Equilibrium: isothermal and isobaric chemical process
For the equilibrium aA + bB cC + dD, under the P,T=const conditions:
dn A =−d ξ a dn B =−d ξ b dn C =d ξ c dn D =d ξ d
dG P, T =0 ∑ i i
µ dn =0 ∑ i iµ ν d ξ=0 νi stoichometrical coefficients (-a,-b,c,d)
since d ξ≠0 due to an equilirium alteration:
o o Δ G oP , T
∑ µi νi =0 {∑ νi µi }=Δ G + ∑ νi RT ln Pi=0
o Δ GP , T =−∑ νi RT ln Pi − =∑ ln ( Piν ) i

P, T
RT
o

Δ G oP ,T −
ΔG
P c
P d P, T

− =ln (∏ Piν ) K P =∏ Pνi =e RT = Ca Db


i i

RT PA PB

Equilibrium: Van't Hoff's law


Δ Go Δ H o ΔS o ∂ ln K Δ H o Δ H o <0 T ↓ K ↑
ln K =−
RT
=−
RT
+
R ∂T
=
RT 2 { Δ H o >0 T ↑ K ↑
It can be used to calculate the value of ∆Ho from two values of the equilibrium constant in a small temperature
range.

Equilibrium: Le Chatelier
The change in one of the variables that determine the balance of a system causes a displacement in the
direction that counteracts said variation:

Δ G=f ( P , T , ξ ) d ( ΔG )= ( ∂ (∂ΔPG) ) dP+ ( ∂ (∂ΔGT ) ) dT + ( ∂ (∂ΔGξ ) ) dξ


( ∂ (∂ΔGP ) )
{
P,T
=ΔV
T, ξ P,ξ T ,ξ
∂2 G ∂2 G
d ( ΔG ) =−Δ S dT + ΔV dP+ ( ) ∂ ξ2 P,T
dξ (equilibrium: ΔG=0, ( ) ∂ ξ2 P ,T
>0 )
( ∂ (∂ΔGT ) ) P, ξ
=−Δ S

ΔH ∂2 G ∂2 G ΔH
0=−
T
dT +Δ V dP+
∂ ξ2 ( ) ( ) P,T

∂ ξ2 P,T
d ξ=
T
d T−Δ V d P ( ∂∂Gξ )P ,T
=Δ G


( )
dT P
=
T
ΔH
∂2 G
(
∂ ξ2 ) {
P ,T
Δ H >0

Δ H <0
( ddTξ ) >0
( ddTξ ) <0
P

P
Δ T ↑=Δ ξ↑

Δ T ↑=Δ ξ↓ | ( ) dξ
dP T
−ΔV
= 2
∂G
( )
∂ ξ2 P , T
Δ V <0 {
Δ V >0 ( ddPξ ) <0
T

( ddPξ ) >0
T
Δ P ↑=Δ ξ↓

Δ P ↑=Δ ξ↑

9
Phase equilibrium
Now it is the same component in equilibrium (P,T=const) between two phases A,B:
dG P , T =0=∑ µi dni µ A + dn A +µB dnB =0 dn A =−dnB (the moles from one phase transfers into the other)
µ A =µ B (thermodynamic condition between phase equilibrium)

Clapeyron equation
From the Gibbs-Duhem equation, if the chemical potentials of a substance are equal in both phases, their
differentials will also be equal:
dU =T dS−P dV +µ dn+ { S dT −V dP+n d µ }=0 dµ=− S̄ dT + υdP − S̄ A dT +υ A dP=− S̄ B dT + υB dP
dP S̄ − S̄ dP Δ H̄
= B A =
dT υB −υ A dT T Δ υ
where we make use of the molar enthalpy of phase change A →B, and the increase in molar volume between
the phases.

Clausius-Clapeyron equation
For liquid-gas systems, we introduce two approximations to the Clapeyron equation:
RT dP P Pf dP Δ H̄ T dT Pf Δ H̄ 1 1
υgas ≫ υliquid Δ υ∼υgas =
P dT
=Δ H̄
R T2
∫P o P
= ∫
R T T2
f

o
ln ( )
Po
= −
R T o Tf( )
Electrical work
Under equilibrium conditions (P,T=const), the Gibbs free energy depends only on mass transfer (open system)
and is equal to non-mechanical (electrical) work. The amount of charge that is transferred in the form of a
mole of electrons times the potential difference:
d GP , T =d ξ ∑ νi µ i=d ξ Δ G P, T =δW rev δ W rev =−νe q e N A d ξ E Δ G P ,T =−ν e F E
o
result extendable to the reference system ∆G .

Osmotic pressure
Consider a system formed by two chambers separated by a
membrane permeable only to the solvent (species A). In one of them
there is only pure solvent (A), while in the other there is a solution
(A+B). Spontaneously, the solvent will move from one chamber to
another so that there will be an increase in pressure in the chamber
with the dissolution: ∆P = Π

The chemical potential of the solvent will be the same in both


chambers.

µ ( P ) = µ o ( P ) ≡ µ ( P+Π , x A ) µ ( P+Π , x A ) = µ o ( P +Π ) +R T ln x A µ o ( P )= µ o ( P+Π ) + R T ln x A


P +Π
V
dF T =−P dV + µ dn+ [−V dP+n d µ ]=0 dµ=
n
o o
dP= µ ( P +Π )− µ ( P ) ∫ Vn dP=− R T ln x
P
A

V 2 n
Π=−RT ln ( 1−x B ) ≈ RT x B Π=c B R T Π= RT c B ( 1+B 2 c B +B3 c B+…+Bn+1 c B )
nT
where we have expressed the chemical potential using the molar fraction (xA + xB = 1), and defined the one for the
pure solvent as its reference value. In addition, ln(1-x) ~ -x for small values of x. The different Bi parameters are
known as the virial osmotic coefficients, and account for the non-ideal bevaviour.

10
The molecular weight obtained from osmometry (Mr) is the number average molar mass:
5
h R T · 10
=
c ρ g Mr
+B c h ( cm ) c
( cmg ) ρ ( cmg )
3 3
R: 8.314 ( molJ K ) g :9.81 kg
( ms ) 2
Mr ( molg )
Activity coefficients: Debye-Hückel
Previously, when defining the chemical potential we used the concept of activity, but this is not correct in the case
of electrolyte solutions since the inter-ionic interactions cause the system to deviate from the ideal behaviour, and
it is necessary to correct it by using coefficients of activity (γi):
2
−B z i √ s 1
µ i= µ oi + RT ln ai ai = γ i c i log γ i= s= ∑ c i z 2i
1 + A ri √ s 2
where A and B are constants which depend on the temperature and the dielectric constant of the solvent, zi
represents the valence of the ion, ri is its hydrodynamic radius, and s is the ionic strength of the solution. For
partially diluted solutions, we can approximate the denominator (Debye-Hückel limit law):
2
log γ i ≈−B zi √s
From an experimental point of view, we cannot determine absolute magnitudes for a single type of ion. In this
case, we use the geometric mean:
1 x y
zA zB [ A ] [ B ] x+ y
Ax B y ⇄ x A + y B γ ±=( γ xA · γ By ) x + y log γ ±=−|z A||z B| B √ s K eq = γ
[ A x By ] ±
B ~ 0.509 for water

Ionic conductivity
An electrolyte is that species that when dissolved in water increases the conductivity of the solution. They can be
strong or weak:

l 1
Λm = κ
o o o o
κ= Λ m=Λ m− A √ c Λm = υ + λ + + υ − λ −
aR c
where R is the resistance of the solution (Ω), l is the distance between
the plates conforming the electrode (cm) and a represents their surface
(cm2); thus κ is the specific conductivity (Ω-1·cm-1), Λm is the molar
conductivity (which decreases rapidly with increasing concentration for
the case of weak electrolytes). This one is related to the limiting molar
conductivity (Λom) for strong electrolytes through the Kohlrausch’s law,
which expression depends on the limiting molar conductivity of each ion
( λ o+ / −), as well as their proportion ( υ + / − ).

The decrease in the case of strong electrolytes with increasing concentration is due to the mutual interference of
the ions in their migrations towards the electrodes.

11
Formal Kinetics
From an empirical point of view, many reactions fulfil that:
r = k [A]α [B]β …
where k is the reaction coefficient (with the necessary units so that speed can be expressed as concentration/time);
A, B, ... are the reacting species; α, β, … are the orders of the reaction with respect to the reactants and their sum
(n) provides the global order of the reaction: n=α+β+…
Molecularity is the number of molecules participating in each fundamental act of the reaction process. Usually it is
equal to the (global) order of the reaction. Partial orders do not always have to be integers:
CH 3 CHO→CH 4 +CO r =k [CH 3 CHO]1,5
In these cases we can think about partial orders, but never for a global one (n must be integer).

Reaction speed
It can be defined unequivocally from the degree of progress of the reaction:
1 dξ 1 d[X]
r= dn x= ν x d ξ r= ν
V dt x dt

Determination of Speed laws (homogeneous systems)


Initial Speeds
Different experiments are prepared with different concentrations and the rates are measured at the beginning of
the reaction, so that almost nothing has reacted and the [X] can be approximated to [X]o:
r o ≈k [ A ]αo [B ]βo ... ln r o=ln k +α ln [ A ]o +βln [B ]o +.. .
In this way, a system of linear equations is proposed. One of the limitations of the method is that complex reactions
(high molecularity) can lead to erroneous interpretations.

Integral Method
1st order:
A → products
1 dn A d [ A] t=0 [ A ]=[ A ]o [ A ] d[ A ] t [ A ]o

vol dt
=−
dt
=k [ A ] { t [A]
−∫[ A ]
[ A]
o
=k ∫0 dt ln
[ A]( )
=k t [ A ]=[ A ]o e −k t

A → P ao
dx x dx t
ao
ao −x x
dt
=k (a o− x) ∫ 0 (ao −x)
=k ∫0 dt ln (
(ao −x ) )
=k t x=a o [ 1−e−k t ]

2nd order:
A 2 → products
d [ A] [A ] d [ A ] t 1 1
− =k [ A ]2 −∫[A ] =k ∫0 dt − =k t
dt o
[ A] 2
[ A ] [ A ]o
dx x dx t x
=k (ao −x )2 ∫0 2 =k ∫0 dt =k t
dt (a o− x) ao (a 0−x )

dx
A
ao
ao −x
+ B
bo
b o− x
→ P

x {dt
=k (ao −x )(b o −x)

∫o (a −xdx
x

o )(bo −x)
=kt
1
=
1
( 1
)( −
1
(a o− x)(bo −x ) b o −a o a o− x b o− x )

12
ao b −x ao (bo −x )
1
∫0 a dx dx 1
)[ ( ) ( )]
x x

( bo −a o )[ −x ∫0 b −x

o o
]( =
b o−ao
ln
ao −x
+ln o
bo
=kt (bo −ao )kt=ln
( bo (ao −x ) )
For second order, you can resort to make identical both initial concentrations, thus reducing to the first case:
dx ao ≡bo dx 2 1 1
=k ( a o −x )( b o −x ) =k ( ao − x ) − =k t
dt → dt ao − x ao

nth order:
An → products
d [ A] [ A] d[ A ] 1 1
− =k [ A ]n −∫[ A ] =kt − =(n−1)k t
dt o
[ A ]n [A]n−1
[ A ]n−1
o

An alternative approach is the isolation method, in which one of the components is introduced in a large excess
and its concentration can be considered to remain constant:
dx
dx
dt
α
=k (a o−x ) (b o−x )
β

so that in a second experience the conditions are reversed...


{
b o ↑↑

ao ↑↑
dt
dx
dt
≈k ' (a o−x )α

≈k ' ' (bo −x)


β
k '=k bβo
α
k ' '=k ao

Half-life Time
Based on the integral method, it defines the time necessary for the concentration of a substance to be reduced by
a half: t½

1st order: ln ( [[AA]] )=k t


o
ln ([ A ]o
[ A ]o /2 )
=k t 1/2 ln 2=k t 1/2

1 1 2 1 1
2nd order: − =k t − =k t 1/2 =k t 1/2
[ A ] [ A ]o [ A ]o [ A ]o [ A ]o
1 1 2n−1 1 2 n−1−1
nth order: − =(n−1)k t − =(n−1)k t 1/2 =(n−1)k t 1/2
[ A ]n−1 [ A ]n−1
o [ A ]n−1
o [ A ]n−1
o [ A ]n−1
o

Equilibriums
In an equilibrium we can consider two opposite reactions, whose speeds are equal when equilibrium is reached:
kd ν A =α
α β π ρ k d [ P ]π [R ]ρ ν B=β kd
A +B ⇄ P+ R r d =k d [ A ] [B ] ri =k i [P ] [ R] r d =r i = K eq =
k i [ A ] [B ]
α β
ν P =π ki
ki ν R=ρ
The equilibrium constant can be related to the kinetic constants in those cases where the stoichiometric
coefficients and the partial orders of the different species coincide:
k1 d [B ]
A⇄B [ A ]=(ao −x) [B ]=x r= =k1 [ A ]−k−1 [B ]=k 1 (ao− x)−k−1 x
dt
k−1
in the dx (ao −xeq ) dx (a0− xeq ) ao x eq
equilibrium dt x ( )
=0 k−1 =k 1
eq
x eq dt
=k 1 (ao −x)−k 1
x eq
x=k 1 a0−k 1 x−k 1
x eq
x +k 1
x eq
x

k 1 a0 t x eq −x a
∫o (x dx−x) =
x

eq
∫ dt
x eq 0
ln
( xeq )=− 0 k 1 t
x eq

The relaxation time (τ) is the time the system needs to reach a new equilibrium situation when a perturbation is
produced (the following example is suitable for the dissociation of weak acids):

13
initial equil. new equil. d[ A]
k1 k 1 [ B] [C ] [ A ]e [ A ]e −x =−k 1 [ A ]+k −1 [ B] [C ]
A ⇄ B +C K eq = = dt
k−1 [ A] [ B ]e [ B]e +x dx
k−1 − =−k 1 ( [ A ]e− x ) +k−1 ( [ B]e +x )( [C ]e+ x )
[C ]e [ C ]e + x dt
dx dx
=k 1 [ A ]e −k 1 x −k −1 [ B]e [C ]e +k −1 [B ]e x +k −1 [ C ]e x+ ( k−1 x2 )≈0 =− ( k 1 +k−1 [ B ]e +k−1 [C ]e ) x=−k obs x
dt dt
1 1
ln x +c=−k obs t x=e
−k obs t
e−c x=x o e
−k obs t
τ =k 1 +k−1 ( [ B]e +[C ]e ) s ()
Parallel reactions
We can differentiate the cases of competitive (the same reactant gives rise to different products) or concurrent
(different reactants lead to the same product):
k1B B k1
d[A] dx
A ↗ r=− =k 1 [ A ]+k 2 [ A ]=(k 1 +k 2 )[ A ] ↘ A r= =k 1 (b o −x)+ k 2 (c o −x )
↘C dt C ↗ dt
k2 k2

Consecutive reactions
In this case, the product of one is the reactant of the next reaction. We can find in series or in chain (many stages in
series):
A k 1 B k 2 C (serial) A → ... →Z (chain)
→ →
d [ A] d [B ] d [C ]
=−k 1 [ A ] =−k 2 [B ]+k 1 [ A ] =k 2 [B ]
dt dt dt
you have to integrate the equations consecutively given certain conditions (t=0 → [B]o=0 & [C]o=0) and
progressively substitute and integrate:
−k t k [ A ]o − k t − k t
[ A ]=[ A ]o e [B ]= 1 1
( e −e ) 1 2

k 2−k 1

Temperature and Speed


The temperature affects the speed of the reactions, generally being that an increase in temperature is
accompanied by an increase in the speed. Arrhenius obtains an empirical form for this behavior:
Ea

k= A e R T
where Ea is known as the activation energy, and A is defined as the pre-exponential or frequency factor and, in
principle, is independent of temperature.
The pre-exponential factor has statistical considerations, while the activation energy defines the minimum energy
that collisions must reach.

Mechanisms: Steady State


In a reaction in which there is the presence of intermediates, the concentration of these is usually small given that
they are species of generally high reactivity:
k 1 [ A ]o −k t − k t
A k1 B k2 C [ B]= ( e −e ) 1 2

→ → k 2−k 1
Considering that:
• the reactivity of the intermediate is a consequence of the fact that k2 >> k1
• if k2 is large, the exponential will tend to zero
• if k1 is small, the exponential will tend to one for a considerable amount of time

14
k 1 [ A ]o −k t −k t k 1 [ A ]o −k t −k t k 1 [ A ]o d [ B]
[B ]= ( e −e )≈
1 2
( e −e )≈
1 2
=0
k 2 −k 1 k2 k2 dt
after a certain initial period, as long as we are not close to the end of the reaction.
Thus, it is assumed that at some point the concentration of all the intermediates remains approximately constant and
much smaller than that of the reactants and products of the reaction.
k1
A +B ⇔ X k 2 C
k−1 →
d[X] d [C ] d[ X ] k 1 [ A ][ B]
=k 1 [ A ][ B ]−k−1 [ X ]−k 2 [ X ] r= =k 2 [ X ] stationary state to X =0 [ X ]=
dt dt dt k−1 +k 2
d [C ] k k
=k [ A ][B ]= 1 2 [ A ][B ]
dt k−1 +k 2
In a mechanism, the sum of the proposed steps must coincide with the global reaction, and the aim is to express
the variation in concentration of all the species with time and solve the system of differential equations (generally
numerically).

Mechanisms: Limiting Step


In mechanisms with consecutive steps, if one of these is much slower than the others, the speed of the overall
reaction will coincide with the speed of the slow step, called the rate limiting/determining step (RDS).
k1
A +B ⇔ X k 2 C from the stationary state (X = intermediate):
k −1 → k k
r= 1 2 [ A ][B ]
- if the second step is the slow one: k −1 +k 2
[ X] - if the second step is the slow one: k 2 ≪k− 1
r=k 2 [ X ] Kc= r=k 2 K c [ A ][ B ]
[ A ][ B ] k1
- if the first step is the slow one: r=k 2 [ A ][B ]
k −1
d[X ] - if the first step is the slow one: k 2 ≫k −1
r= =k 1 [ A ][ B]−k −1 [ X ]≈k 1[ A ][ B ]
dt
r=k 1 [ A ][ B ]
in equilibrium there is no return, since X does not accumulate

Transition State Theory


Transition states are high-energy species which are found in between stable ones (or semi-stable such as
intermediates):
R ⇄ X ‡→ P [ X‡ ]=K ‡ [ R]

d [P] k T k T k B T − ∆G RT
n− 1

dt
=[ X ‡] B = B K ‡ [ R ]
h h
k TST =
h
e RT
( )
Po
where h/kB·T represents the average time it takes for “activated” molecules (with enough energy to go from R to P)
to cross the region of the transition state (X‡). This theory is based on two assumptions: a) there is no recrossing
(any molecule that crosses the transition state region will inevitably reach the products), and b) there is no
tunnelling effect (or other quantum nuclear effects).

In addition, for an unimolecular process (n=1):


2 ∂ ln k k B T ∆ H ‡ ∆ S‡ ∂ ln k 1 ∆ H ‡ ‡
E a=R T ln k= ln − + = + E a=n R T +∆ H
∂T h RT R ∂T T R T2
‡ ‡
∆S
Ea k B T ∆ H ‡ ∆ S‡ ∆H

kBT k B T n+ ∆RS
ln k= ln A − ln − + = ln A−1− ln A =ln + ln e+ ln e R A= e
RT h RT R RT h h

15
Since the species I is more exergonic than R, the rate
determining step corresponds to the I→P step (through the T2
transition state). The final reaction energy is always defined
referring to the reactants (∆GR = GP – GR).

Finally, the kinetic isotope effect (KIE) is a phenomenon associated with isotopically substituted molecules
exhibiting different reaction rates. Isotope replacement does not change the electronic structure of the molecule
or the potential energy surfaces of the reactions the molecule may undergo, and only the rate of the reaction is
affected. It is defined as the ratio between the kinetic constant for the light isotope with respect to that of the
heavy one (under the TST):
‡ ‡
∆ G Heavy − ∆G Light
k Light RT
KIE= =e
k Heavy

________________________________________________________________________________________________

k N 2
k j 2 k 2 ∂d
P ( x )= ∑ a j x {x i , yi }N d =∑ ( P ( xi )− y i ) =0
j=0 i=1 ∂ aj

N ∑x ... ∑ xk ao ∑y

( ∑x
...
∑ xk
...
...
∑ x k +1
...
...
...
∑ x k +1 ·
∑ x 2k
...
)( ) ( )
... =
...
ak
∑ x·y
...
∑ xk · y
2
r =1−
d
2

∑ ( y i− ȳ ) 2 =1−
( N −1 )
2
d
2
sdev ( y )

y=m· x m=
∑y
∑x
________________________________________________________________________________________________

Numerical resolution: A → B → C
import numpy
import matplotlib.pyplot as plt

k1 = 0.7
k2 = 2 * k1
a0 = 0.1

dt = 0.001
t = numpy.arange( 10000 ) * dt
a = [ a0 ]
b = [ 0 ]
c = [ 0 ]
for i in range( 1, t.shape[0] ):
# d[A]/dt = - k1 [A]
rk1 = - k1 * a[i-1]
rk2 = - k1 * ( a[i-1] + 0.5 * rk1 * dt )
rk3 = - k1 * ( a[i-1] + 0.5 * rk2 * dt )
rk4 = - k1 * ( a[i-1] + rk3 * dt )
a.append( a[i-1] + dt * ( rk1 + 2 * rk2 + 2 * rk3 + rk4 ) / 6 )
# d[B]/dt = k1 [A] - k2 [B]
rk1 = k1 * a[i-1] - k2 * b[i-1]
rk2 = k1 * a[i-1] - k2 * ( b[i-1] + 0.5 * rk1 * dt )
rk3 = k1 * a[i-1] - k2 * ( b[i-1] + 0.5 * rk2 * dt )
rk4 = k1 * a[i-1] - k2 * ( b[i-1] + rk3 * dt )
b.append( b[i-1] + dt * ( rk1 + 2 * rk2 + 2 * rk3 + rk4 ) / 6 )
# d[C]/dt = k2 [B]
c.append( c[i-1] + dt * k2 * b[i-1] )

plt.grid( True )
plt.plot( t, a, '-', label = "[A]" )
plt.plot( t, b, '-', label = "[B]" )
plt.plot( t, c, '-', label = "[C]" )
plt.legend( loc = "center right", fontsize = "small" )

16
Surface Tension
Whenever there are two phases in contact, phenomena associated with the existence of a region between both
phases (or interface) always appear. In a gas/liquid system, the solvent molecules are affected by the thermal
agitation, and the interaction with the other surrounding molecules results in a null force. On the other hand, the
molecules of the interface, because the interaction with the molecules of the gas phase is less strong, will develop
a non-zero force directed toward the core of the solvent, which is the final responsible for the surface to be
minimized (spherical droplets of water: smallest liquid/gas contact surface). This is the reason why we will need to
do work to increase the surface area of the interface.
1 2
4 4 4
V sphere= π r3=L3 =V cube
3
L= ( )
3
π 3
r S sphere=4 π r 2 < S cube =6 ( )r3
π 3 2

The surface tension (γ), is a force per unit length (N/m) and is defined always positive (γ>0). The lower the affinity
of a substance with the phase in contact, the greater the surface tension between them. It depends on the
chemical nature of the substances and the temperature, but not on the contact surface (A).
−3 N
δ W = γ dA γ 20ºC
air /water =75.75 · 10
m
Experimentally, it is observed that the temperature reaches a certain value (critical temperature, T o), at which both
phases begin to mix and the surface tension becomes almost zero. Guggenheim proposed the following empirical
equation (11/9 is fine for many organic liquids):
11
T
γ = γ o 1− ( To ) 9

Young-Laplace equation
Suppose a sphere of liquid with a given radius (r) surrounded by a gas (Pgas), in the equilibrium:
F gas+ F surf = Fliq F gas =P gas 4 π r 2 Fliq =Pliq 4 π r 2 dW= F dr F surf =( γ dA)/ dr= γ 8 π r

Pgas 4 π r 2 + γ 8 π r=Pliq 4 π r 2 Pliq =Pgas +
r
Thermodynamically, we consider two phases (α and β) separated by an interface without volume σ (closed,
reversible and isothermal):
dU =T (dSα +dS β +dS σ )−P α dV α −Pβ dV β + γ dA V T =V α +V β dV T =0 dV α =−dV β
2 2γ
sphere dV α =4 π r dr
dU =0 , dS T =0 dV α ( Pβ −P α ) + γ dA=0
dA=8 π r dr {
P α −P β =
r

Capillarity

In this case, the pressure difference


responsible of the liquid rise, has its
source in the gas phase: Pgas > Pliquid, being
R the capillary radius, and r the radius of
the formed meniscus.

2γ R 2 γ cos θ
Pgas−Pliquid = =g h ( ρ liquid − ρ gas ) cos θ = h=
r r g ( ρ liquid− ρ gas ) R
Values of θ lesser than 90º (as in the picture) lead to a rise of liquid inside the capillary (capillary wetting/soaking),
meanwhile larger values (>90º) produce a descent of the liquid (such as the case of liquid Hg).

17
The exact treatment of capillary rise must take into account the deviation of the meniscus from sphericity.

Surface concentration
Given an open system composed by two phases (α and β) separated by one interface (σ), the Gibbs approach is
based on the assumption that the interface has no volume (Vσ=0, as opposed to the Guggenheim approach for
which Vσ≠0):
µ i = µ i, α = µ i, β = µ i , σ n i=ni , α +ni , β + ni, σ dU σ =T dSσ + γ dA+∑ µ i ni, σ
where the chemical potential is the same for a given species in all phases, and the species are distributed
throughout all phases. Thus, the application of Gibbs-Duhem under isothermal conditions, leads to:
ni , σ
A d γ +∑ n i, σ d µ i =0
{
Γ i, σ =
A
surface concentration

d µ i= R T d ln c i , α =R T
dc i, α
ci ,α
( dcd γ )=−Γ
i, α

where the surface concentration accounts for the amount of species i which reaches the interface (ni,σ), arising
i ,σ
RT
c i, α

from an initial dissolution in the α phase (ci,α). This equation establishes how the surface tension between two
phases is modified/affected, by the simple addition of some chemical species to one of the phases (α), as a result of
a migration of this species towards the interface.

The Pockels’ point is found when the surface reaches its maximum of concentration
(of species i), and no further change of γ can be achieved.

Kelvin equation
Recalling Gibbs-Duhem for an open isothermal process, without chemical reaction:

n dµ=V dP d µi =υi dPi d µ gas =dµ liquid υliquid dPliquid =υgas dPgas P liquid −Pgas =
r
dPliquid −dP gas=d ( 2rγ ) υgas 2γ
υliquid dPgas −dPgas =d r
υ gas−υliquid
( ) 2γ
υliquid dPgas =d r ( )
RT
υgas ≫υliquid υliquid≈const υgas =
P gas
1 RT 2γ P υliquid 2 γ
υliquid ∫ P dP gas =∫ d r
gas
( )
ln =
Po R T r
where Po is the standard vapour pressure of the liquid at given the temperature. As a result, the vapour pressure of
a spherical droplet (with radius r) will be larger than the standard value, and therefore the greater its tendency to
evaporate (the vapour pressure is the pressure needed to apply to a liquid-gas system in order to avoid molecules
of the liquid going to the gas phase, thus the higher the vapour pressure the faster a liquid evaporates).
In a capillary, since the liquid has a curvature with a negative radius, it is common to observe condensation on the
internal walls (this explanation can also be applied to cavitation within a liquid).

Cohesion, Adhesion and Detergency


Adhesion is defined as the change in free energy per unit area necessary to separate two media, A and B, from
contact to infinity.

18
A| B → A … B ∆ Gadhesion = γ A + γ B + γ A , B
Since it is defined as a work of separation, the condition for both phases to be stuck together in contact (A|B) is
that ∆Gadhesion > 0. In the case of cohesion, we come across that A=B, so ∆G cohesion = 2 γA, which explains why liquids
tend to join together.
In the case of detergency, we find a system made up of three components: textile, fat and water.
Textile | Fat … Water ∆G detergency = γ W , F + γ W , T− γ T , F
where now we look for ∆Gdetergency < 0. Most detergents/soaps operate by decreasing γWater,Fat . Finally, the presence
of foam favours the suspension of fat, but does not directly affect to the value of ∆Gdetergency.

Adsorption
Adsorption is a phenomenon of fixation of particles on a surface. We find two kinds:
• Physisorption (or physical adsorption): the molecules are fixed to the surface through weak interactions
(usually van der Waals, or dipolar interaction), without giving rise to the formation of chemical bonds
between the particles and the surface. Generally, disorders occur in the geometry of the molecules when
adsorption occurs (bond elongation, ...), and it is a reversible process (but we will have to provide energy
for desorption to occur).
• Chemisorption: is a selective process and results in the formation of chemical bonds between the
molecules and the surface (generally covalent). It only takes place in a monolayer (the next layer(s) will be
due to physisorption), and we cannot ensure that it is a reversible process. The ability of surfaces to
promote bond dissociation is the critical part of understanding why surfaces can catalyse reactions.

A key concept is the fractional occupancy/surface coverage (θ), defined as the ratio between surface positions that
are occupied, with respect to the total available positions:
occupied positions
θ=
available positions

Langmuir adsorption isotherm


It is based on 3 assumptions:
• No adsorption after an initial monolayer
• All positions on the surface of the solid are equivalent (∆Hadsorption of each binding site is the same)
• Neighbouring positions that are occupied do not affect the adsorption of unoccupied ones (there are no
interactions between adsorbed molecules)
total sites: N
A ( g)+M (surf )⇄ AM (surf)

dt
=k adsorption P free_sites−k desorption occupied_sites
{
occupied_sites : N θ
free_sites : N ( 1−θ )
at the equilibrium both adsorption and desorption become equal, thus dθ/dt = 0
k
K eq = adsoprtion k adsoprtion P N ( 1− θ )=k desorption N θ K eq P−K eq P θ =θ K eq P= θ ( 1+ K eq P )
k desorption
K P 1 1 1 ∂ ln K eq ∆ H oadsorption
θ = eq θ = K eq P +1 =
1+ K eq P ∂T R T2
[10.1111/j.1472-8206.2008.00633.x] Hill equation
In spite of these assumptions, the Langmuir adsorption isotherm is very useful for many systems at low pressures
where only monolayer adsorption is involved. On the other hand, increasing the pressure makes it essential to
consider physisorption in multiple layers, and thus relying on the Brunauer, Emmett and Teller (BET) adsorption
isotherm.

19
BET adsorption isotherm
In this multilayer adsorption, it is assumed that the surface possesses uniform, localized sites and that adsorption
on one site does not affect adsorption on neighbouring sites (just as before). It is further assumed that molecules
can be adsorbed in second, third,..., and nth layers, with the surface area available for the nth layer equal to the
coverage of the (n-1)th layer:

2 3 ∞
θ o=

θ 1=
10
3
10
θ 2=

θ 3=
10
2
10
} N =10
(available positions) θ T=
∑i=0 θ i =1
N particles
N

=∑i=1 i θ i

Considering the equilibrium expression of each adsorption layer:


θ1
A (g)+M (surf)⇄ AM (surf ) K 1= θ 1= K P θ o
θ oP
1
A (g)+ AM (surf)⇄ A2 M (surf) θ 2=K 2 P θ 1 =K condensation K θ o P2 K cond =
Po
θ 2 =K θ o P
( PP )
o

P K P K P
z= θ i =K P θ o z i−1 c= K P= K P o = =c z θ i =c θ o z i
Po K cond P o K cond P o
where the first equilibrium constant (K1 or K) will be higher than the remaining ones (K 2, …), since the adsorption
that occurs directly on the surface will be stronger than those of the subsequent layers. In fact, the rest of the
equilibrium constants are assumed to have practically the same value (K2 ≈ K3 ≈ ...), and equal to the gas
condensation equilibrium constant (Kcondensation), which is defined as the inverse of the liquid vapour pressure (Po).
∞ ∞ ∞ z
N part ∑i=1 i θ i
N
= ∞ =
c θ o ∑i=1 i z i

∑i=0 θ i θ o +c θ o ∑i=1 zi
In addition, the following relationships can be applied:
{∑i =1 i z i= ( z−1)
∞ −z
∑i=1 z i= (z−1)
2 N part
N
=
c z /( z −1)2
1−c z /( z −1)
N part
N
=
cz
( 1− z ) ( 1+ c z − z )

∆ H o condensation−∆ H o adsorption ∆ H o desorption− ∆ H o vaporization


N particles
c≈e RT
≈e RT
N
=
Volume
Volume monolayer ( NN )
part

c ≫1

1
1− z

Finally, other isotherms, such as the Freundlich isotherm, attempt to incorporate the effects of the interactions
between substrates when they are adsorbed on the surface, in spite of using a simple equation:
θ = c1 Pc 2

_____________________
The Langmuir model (a) is based on the assumption that only a monomolecular layer of noninteracting solutes is
formed on the sorbent surface, with all adsorption sites being of equal adsorption energy. The Freundlich (c)
model is used to describe monolayer adsorption on heterogeneous surfaces with adsorption sites having differing
adsorption energies and rates of adsorption. The BET model (b) provides for multilayered adsorption on different
areas of the surface, where it is not required for one layer to be completed before another layer begins to form on
top of it.

[10.1002/jbm.a.35235]

20
Catalysis
A catalyst modifies the rate of conversion of reactants into products, recovering at the end of the reaction (it does
not modify the thermodynamics, only the kinetics).
It generally acts by decreasing (or smoothing) the potential energy curve, so that the transition state forms a
complex with the catalyst that decreases its energy; but it can also act by modifying the mechanism through which
the reaction takes place.

Homogeneous catalysis
The catalyst and the reactants are in the same phase (normally in solution). The main problem is usually recovering
the catalyst. Regarding the mechanism of action, we can propose the following one in which an activated species
between one of the reactants and the catalyst is formed (ACat), to which the steady state approximation is applied:
k+ 1 k2
A +B+Cat → P+ Cat A+Cat ⇄ ACat+ B → Cat + P
k −1
d [ ACat ] k+ 1 [ A][Cat ]
≈0=k + 1 [ A][Cat ]−k−1 [ ACat ]−k 2 [ ACat ][ B] [ ACat ]=
dt k−1 +k 2 [B]

k 2 [B]≫ k−1 r =k+ 1 [Cat ] [ A ]=k ' [ A]


r=
d [ P]
dt
=k 2 [B][ ACat ] r=
k2 k +1 [ A][B][Cat ]
k −1 + k 2 [B]
{k 2 [B]≪ k−1 r=
k +1 k2
k−1
[Cat ] [ A][B]=k ' ' [ A][B]

Acid-base Catalysis
General catalysis is detected by rate measurements at constant pH but different buffer concentrations: if the rate
of the reaction does increase with increasing concentration of the buffer at a constant pH, the rate of the reaction
is thus shown to be dependent upon the concentration of a component of the buffer (HA in the case of acid
catalysis).
On the other hand, if the rate of the reaction is independent of the concentration of the buffer, Specific acid
catalysis is involved; since the rate specifically depends on the concentration of hydronium ion (H3O+).
Thus, we should consider three contributions to the overall chemical transformation:
r=k o [ S ]+k H O [ H 3 O+ ][ S ]+k HA [ HA ][ S ]
+ r =k obs [ S ] k obs=k o +k H O [ H 3 O+ ]+k HA [ HA ]
+
3 3

where ko accounts for the kinetics of the uncatalyzed reaction, and the remaining ones are for the specific and
general acid catalysis.

In specific acid catalysis a reactant is in equilibrium with its conjugate acid, which position of the equilibrium is
determined by the pH of the solution, and the proton transfer is not rate-determining step. Conversely, in general
catalysis the proton transfer is part of the rate-determining step.

21
Heterogeneous catalysis
The main characteristic is that the catalyst is insoluble in the reaction medium. Catalysts where the surface/mass
ratio is large will be of interest.
Taylor's Law: for a substance to act as a heterogeneous catalyst, at least one of the reactants must be chemically
adsorbed (chemisorption) on its surface.
Sabatier principle: a balance must be reached between activity and adsorption force.

Steps of the kinetic mechanisms


1) diffusion of reactants to the surface of the catalyst
2) adsorption (chemisorption, exothermic process: bond are formed)
3) chemical reaction on the surface
4) desorption of surface products (endothermic process)
5) diffusion of products
Any of the five steps can control the speed of the process

Enzymatic catalysis
An enzyme is a biological catalyst of protein nature that has a region where catalysis takes place (active site). There
are enzymes that require additional molecules in addition to the substrate, which we call cofactors (or coenzymes
if they are organic in nature):
apoenzyme + cofactor → holoenzyme
If the cofactor is modified during the catalysed reaction, it must be regenerated in some subsequent reaction.

Briggs-Haldane
Stationary state approximation on the enzyme-substrate complex:
k +1 k2 d [ ES] [ E][ S ] k −1 +k 2
E+S ⇄ ES → E +P =0=k +1 [ E ][ S]−k −1 [ ES ]−k 2 [ ES ] = =K M [ E ]o =[ E]+[ ES]
dt [ ES ] k +1
k−1

[ E]o −[ ES ] [ E]o [ S] [ E]o [ S] r max [ S ]


[S ]= K M [ ES ]= r=k 2 [ ES]=k 2 [ S ]↑↑ ; r max=k 2 [ E ]o r=
[ ES ] [S ]+ K M [S ]+K M [ S ]+K M
r max k cat k cat
k cat = [S ]≪ K M r≈ [ E]o [ S] catalytic efficiency=
[ E ]o KM KM
They found the same expression as Michaelis-Menten, being the difference in the definition of the K M, which in
the original derivation only considered the inverse of the initial equilibrium constant (k-1/k+1).

Lineweaver-Burk
1 [ S]+K M 1 KM 1 1
= = +
r r max [ S ] r r max [ S] r max
There are other expressions derived from this one, but all of them experience problems (deviations from linearity)
at low substrate concentrations.

22
Eisenthal-Cornish-Bowden
1 [S ]+K M r max K M r
= = +1 r max= K +r
r r max [ S ] r [S] [S] M

Inhibition
Competitive: Inhibitor and reactant compete for the active site
E+S+I ⇄ ES → E+P [ E][ S ] [ E ][ I ] [I ]
K M= K I= α =1+ [ E]o =[ E]+[ ES ]+[ EI ]
E+S+I ⇄ EI [ ES ] [ EI ] KI
[I ] [ ES ] [ I] [ S ]+ α K M r max [ S ]
(
[ E ]o =[ ES ]+[ E ] 1+
KI )
=[ ES]+K M
[S] (
1+
KI )
=[ ES ] ( [S] ) r=
[ S]+ α K M

Uncompetitive: Inhibitor binds in a different site, once the enzyme-substrate complex is formed
E+S ⇄ ES → E+P [ E ][ S] [ ES ][ I ] [I]
K M= K I '= α '=1+ [ E]o =[ E ]+[ ES ]+[ ESI ]
E+S ⇄ ES+I ⇄ ESI [ ES] [ ESI ] KI'
K [I ] α ' [ S ]+K M r [ S]
[ E ]o= M [ ES ]+[ ES ]+
[ S] KI '
[ ES]=[ ES] ( [S] )r= max
α ' [S ]+K M

Mixed: Inhibitor binds in a different site, both on the enzyme and on the enzyme-substrate complex
E+S ⇄ ES → E+P r max [ S]
E+I ⇄ EI [ E ]o =[ E]+[ ES]+[ EI ]+[ ESI ] r=
α ' [ S ]+ α K M
E+S ⇄ ES+I ⇄ ESI
The special case when α = α’, has been classically defined as noncompetitive inhibition.

Finally, irreversible inhibitors bind covalently with or destroy a functional group on an enzyme that is essential for
the enzyme’s activity, or form a particularly stable non-covalent association. In those cases:
r max =k cat ( [ E ]o −[ I ] )

23

You might also like