5
5
• w = Force * distance
• In an experiment of chemical reaction (a container with
piston) -
• Piston will move if the system works on it
Pressure exerted by the system on the surface of piston
Pressure by the system = Force/Area
Force= P * Area
Therefore, w =P * Area * the distance the piston has moved (
let’s say x)
X= P * A * x
W= P* Δ V
Δ V implies the change in volume that happened
Internal Energy
• Source:https://chem.libretexts.org/Bookshelves/General_Chemistry/Map%3A_Chemistry_-_The_Central_Science_(Brown_et_al.)/19%3A_Che
mical_Thermodynamics/19.2%3A_Entropy_and_the_Second_Law_of_Thermodynamics
Therefore, our experiences, tell us that there is a
preferred direction to many natural processes.
Like- a cup of coffee gradually loses heat to its
surroundings as it cools, or when the ice in a glass of
lemonade absorbs heat as it melts. Instead We
would be surprised if a cup of coffee suddenly grew
hotter until it boiled or the water in a glass of
lemonade froze on a hot summer day, even though
neither process violates the first law of
thermodynamics.
Many chemical and physical processes are
reversible and yet tend to proceed in a direction
in which they are said to be spontaneous. This
raises an obvious question: What makes a
reaction spontaneous? What drives the
reaction in one direction and not the other?
So many spontaneous reactions are exothermic that
it is tempting to assume that one of the driving
forces that determines whether a reaction is
spontaneous is a tendency to give off energy.
BUT
There are also spontaneous reactions, however, that
absorb energy from their surroundings. At 100oC,
water boils spontaneously even though the reaction
is endothermic.
H2O(l) → H2O(g) ΔHo = 40.88 kJ/mol
• Thus, the tendency of a spontaneous reaction
to give off energy can't be the only driving
force behind a chemical reaction. There must
be another factor that helps determine
whether a reaction is spontaneous. This
factor, known as entropy, is a measure of the
disorder of the system.
Entropy and the Second Law of
Thermodynamics
• The second law of thermodynamics describes the
relationship between entropy and the spontaneity of
natural processes.
• Second Law: In an isolated system, natural processes
are spontaneous when they lead to an increase in
disorder, or entropy.
• This statement is restricted to isolated systems to
avoid having to worry about whether the reaction is
exothermic or endothermic. By definition, neither heat
nor work can be transferred between an isolated
system and its surroundings.
• We can apply the second law of
thermodynamics to chemical reactions by
noting that the entropy of a system is a state
function that is directly proportional to the
disorder of the system.
• ΔSsys > 0 implies that the system becomes
more disordered during the reaction.
• ΔSsys < 0 implies that the system becomes
less disordered during the reaction.
• For an isolated system, any process that leads
to an increase in the disorder of the system
will be spontaneous. The following
generalizations can help us decide when a
chemical reaction leads to an increase in the
disorder of the system.
• Solids have a much more regular structure than
liquids. Liquids are therefore more disordered
than solids.
• The particles in a gas are in a state of constant,
random motion. Gases are therefore more
disordered than the corresponding liquids.
• Any process that increases the number of
particles in the system increases the amount of
disorder.
How H and S affect direction of
reaction
• The sign of ΔH for a chemical reaction affects
the direction in which the reaction occurs.
Spontaneous reactions often, but not always,
give off energy.
• The sign of Δ S for a reaction can also
determine the direction of the reaction.
In an isolated system, chemical reactions occur
in the direction that leads to an increase in the
disorder of the system.
• In order to decide whether a reaction is
spontaneous, it is therefore important to
consider the effect of changes in both
enthalpy and entropy that occur during the
reaction.
The Third Law of Thermodynamics
• For any element at any particular temperature, we define the standard enthalpy of formation to
be zero. When we define standard enthalpies of formation, we choose the elements in their
standard states as a common reference state for the enthalpies of all substances at a given
temperature. While we could choose any arbitrary value for the enthalpy of an element in its
standard state, choosing it to be zero is particularly convenient
• The enthalpy data are given in terms of the
standard-state enthalpy of formation of each
substance, Δ Hfo.
• This quantity is the heat given off or absorbed
when the substance is made from its elements in
their most thermodynamically stable state at
0.1 Mpa (1 bar).
• The enthalpy of formation of AlCl3, for example, is
the heat given off in the following reaction.
standard conditions:
• 1 atm for all gases and a concentration of 1.0 M (1 mol/L) for all species in
solution.
• Each pure substance must be in its standard state, which is usually its most
stable form at a pressure of 1 atm at a specified temperature.
• Enthalpies of formation measured under these conditions are
called standard enthalpies of formation (ΔHof)
• The enthalpy change for the formation of 1 mol of a compound from its
component elements when the component elements are each in their
standard states.
• The standard enthalpy of formation of any element in its most stable
form is zero by definition. (which is pronounced “delta H eff naught”).
• The standard enthalpy of formation of any element in its standard state is
zero by definition.
Example-oxygen can exist as ozone (O3), atomic oxygen (O), and molecular oxygen (O2), O2 is
the most stable form at 1 atm pressure and 25°C. Similarly, hydrogen is H 2(g), not atomic
hydrogen (H). Graphite and diamond are both forms of elemental carbon, but because graphite
is more stable at 1 atm pressure and 25°C, the standard state of carbon is graphite .
Therefore, O2(g), H2(g), and graphite have ΔHof values of zero.
Source:
https://chem.libretexts.org/Courses/University_of_Arkansas_Cossatot/UAC%3A_Chem_1024/05._Thermochemistry/5.7%3A_Enthalpie
s_of_Formation
• The standard enthalpy of formation, ∆Hfo, is the
change
in enthalpy when one mole of a substance is formed
from its elements under a standard pressure of 1
atm.
• The heat of formation of any element in its
standard state is defined as zero.
• The standard enthalpy of reaction, ∆Hfo, is the
sum of the enthalpy of the products minus the
sum of the enthalpy of the reactants.
• ∆Hfo =
• The enthalpy data in this table are therefore relative
numbers, which compare each compound with its
elements.
• Enthalpy data are listed as relative measurements
because there is no absolute zero on the enthalpy
scale. All we can measure is the heat given off or
absorbed by a reaction. Thus, all we can determine is
the difference between the enthalpies of the
reactants and the products of a reaction. We
therefore define the enthalpy of formation of the
elements in their most thermodynamically stable
states as zero and report all compounds as either
more or less stable than their elements.
• Between Graphite and Diamond Which form
of carbon will be considered for standard
enthalpy of formation????
and then the product of that reaction in turn reacts with water to form phosphorus acid
2P2O5+6H2O→4H3PO4……..(2)
In the above equation the P2O5 is an intermediate, and if we add the two equations the intermediate can cancel
out. Hess's law states that if two reactions can be added into a third, the energy of the third is the sum of the energy
of the reactions that were combined to create the third.
• equation 1: P4+5O2→2P2O5………..ΔH1
• equation 2: 2P2O5+6H2O→4H3PO4………….ΔH2
• equation 3 (equation 1 + equation 2): P4+5O2+6H2O→3H3PO4………….ΔH3
Enthalpy is a state function which means the energy change between two states is independent of the path. Since
equation 1 and 2 add to become equation 3, we can say:
• ΔH3=ΔH1+ΔH2(4)
• Hess's Law says that if equations can be combined to form another equation, the enthalpy of reaction of the
resulting equation is the sum of the enthalpies of all the equations that combined to produce it.
Source:
https://chem.libretexts.org/Courses/University_of_Arkansas_Little_Rock/Chem_1402%3A_General_Chemistry_1_(Belford)/Text/5%3A_Energy_and_Chemical_Re
actions/5.7%3A_Enthalpy_Calculations
The additive property of ΔH°rxn values will be applied in this experiment in order to
determine the enthalpy change associated with the
burning of magnesium metal in air.
• For the combustion of magnesium, reaction (1) one possible set of
reactions is:
(2) Mg(s) + 2 H+(aq) → Mg2+(aq) + H2(g) ΔH°rxn(2) = negative
number
(3) MgO(s) + 2 H+(aq) → Mg2+(aq) + H2O (l) ΔH°rxn(3) = negative
number
(4) 1/2 O2 (g) + H2(g) → H2O (l) ΔH°rxn(4) = negative number
Unfavorable, or
ΔG > 0
non-spontaneous reactions:
• Reactions are classified as either exothermic (
ΔH < 0) or endothermic (Δ H > 0) on the basis
of whether they give off or absorb heat.
Reactions can also be classified as exergonic
(Δ G < 0) or endergonic (Δ G > 0) on the basis
of whether the free energy of the system
decreases or increases during the reaction.
• When a reaction is favored by both enthalpy
(Δ H < 0) and entropy (Δ S > 0), there is no
need to calculate the value of Δ G to decide
whether the reaction should proceed. The
same can be said for reactions favored by
neither enthalpy (Δ H > 0) nor entropy (Δ S <
0). Free energy calculations become important
for reactions favored by only one of these
factors.
• The beauty of the equation defining the free
energy of a system is its ability to determine
the relative importance of the enthalpy and
entropy terms as driving forces behind a
particular reaction. The change in the free
energy of the system that occurs during a
reaction measures the balance between the
two driving forces that determine whether a
reaction is spontaneous.
• If the reaction is run at constant temperature,
this equation can be written as follows.
ΔG = Δ H - T Δ S
• The change in the free energy of a system that
occurs during a reaction can be measured under
any set of conditions. If the data are collected
under standard-state conditions, the result is the
standard-state free energy of reaction (ΔGo).
ΔGo = Δ Ho - TΔ So
• What is the standard free-energy change (ΔG°)?
- is the change in free energy when one substance or a set of substances in their
standard states (the state g/l/s in which they are stable at standard condition; temp. not
specified at standard condition)is converted to one or more other substances, also in their
standard states.
The standard free energy of formation (ΔG° f)of a compound is the change in free energy that
occurs when 1 mol of a substance is formed from the component elements in their
standard states.
By definition, the standard free energy of formation of an element in its standard
state is zero at 298.15 K.
Tabulated values of standard free energies of formation allow chemists to calculate the values of ΔG° for a wide variety of
chemical reactions rather than having to measure them in the laboratory.
Source:
https://chem.libretexts.org/Bookshelves/General_Chemistry/Map%3A_General_Chemistry_(Petrucci_et_al.)/19%3A_Spontaneous_Ch
ange%3A_Entropy_and_Gibbs_Energy/19.6%3A_Gibbs_Energy_Change_and_Equilibrium
• standard states (the state g/l/s in which it is stable at 0.1
Mpa/1 M concentration, temp. not specified at standard
condition).
• For water at –10°C, this reaction is H2(g,−10°C, 1 bar)+ 12
O2(g,−10°C, 1 bar) →H2O(s,−10°C, 1 bar)
• For water at +10°C, it is H2(g,+10°C, 1 bar)+ 12 O2(g,+10°C, 1
bar) →H2O(liq,+10 ° C, 1 bar)
• For water at +110 ° C, it is H2(g,+110 ° C, 1 bar)+ 12 O2(g,+110
° C, 1 bar) →H2O(g,+110 ° C, 1 bar)
• .
The Effect of Temperature on the Free
Energy of a Reaction
• The balance between the contributions from the
enthalpy and entropy terms to the free energy of a
reaction depends on the temperature at which the
reaction is run.
• The equation used to define free energy suggests that
the entropy term will become more important as the
temperature changes.
» ΔGo = Δ Ho - T Δ So
• The relationship between the free energy of reaction at any
moment in time (ΔG) and the standard-state free energy of
reaction (ΔGo) is described by the following equation.
ΔG = ΔGo + RT ln Q
In this equation, R is the ideal gas constant in units of J/mol-K,
T is the temperature in kelvin, ln represents a logarithm to the
base e, and Q is the reaction quotient at that moment in time.
• As we have seen, the driving force behind a chemical
reaction is zero (ΔG = 0) when the reaction is at
equilibrium (Q = K).
0 = ΔGo + RT ln K
• We can therefore solve this equation for the relationship
between ΔGo and K.
ΔGo = - RT ln K
This equation allows us to calculate the equilibrium constant
for any reaction from the standard-state free energy of
reaction, or vice versa.
• ΔG = ΔG⁰ + RT ln Q
ΔG⁰ = – RT ln K
We now have a way of relating the equilibrium constant directly to
changes in enthalpy and entropy. As well, we can now determine the
equilibrium constant from thermochemical data tables or determine the
standard free energy change from equilibrium constants.
SOURCE: https://opentextbc.ca/introductorychemistry/chapter/free-energy-under-nonstandard-conditions-2/
The relationship between ΔGo and the equilibrium constant for a
chemical reaction is illustrated by the data in the table below.
Values of ΔGo and K for Common Reactions at 25oC
The Relationship Between Free Energy
and Equilibrium Constants
• The difference between ΔGo and Δ G for a
reaction is important. There is only one value of Δ
Go for a reaction at a given temperature, but
there are an infinite number of possible values of
Δ G.
• The figure below shows the relationship between
Δ G for the following reaction and the logarithm
to the base e of the reaction quotient for the
reaction between N2 and H2 to form NH3.
N2(g) + 3 H2(g) ↔2 NH3(g)
Interpreting Standard-State Free
Energy of Reaction Data
• What does the value of ΔGo tell us about the
following reaction?
N2(g) + 3 H2(g) ↔ 2 NH3(g) ΔGo = -32.96 kJ at 25°C
But at 500 °C ΔGo =61.4 KJ
By definition, the value of ΔGo for a reaction measures the
difference between the free energies of the reactants and
products when all components of the reaction are present
at standard-state conditions.
At which temperature the forward reaction is favourable?
• The sign of ΔGo tells us the direction in which the
reaction has to shift to come to equilibrium.
• The fact that ΔGo is negative for this reaction at 25oC
means that a system under standard-state conditions
at this temperature would have to shift to the right,
converting some of the reactants into products, before
it can reach equilibrium. The magnitude of ΔGo for a
reaction tells us how far the standard state is from
equilibrium. The larger the value of ΔGo , the further
the reaction has to go to get to from the
standard-state conditions to equilibrium.
• Assume, for example, that we start with the following
reaction under standard-state conditions, as shown in
the figure below.
• N2(g) + 3 H2(g) ↔ 2 NH3(g)
• At 25°C ∆Go therefore
describes this reaction
only when
all three components
are present at 1 atm pressure
(as the reactant-product are in
standard state)
Standard-State Free Energy of
Reaction
• The value of ΔG at that moment in time will be equal
to the standard-state free energy for this reaction,
ΔGo.
When Q = 1 (measured from partial pressure)
Δ G = Δ Go (ΔG = ΔG⁰ + RT ln Q)
• As the reaction gradually shifts to the right, converting
N2 and H2 into NH3, the value of Δ G for the reaction
will decrease. If we could find some way to harness
the tendency of this reaction to come to equilibrium,
we could get the reaction to do work. The free energy
of a reaction at any moment in time is therefore said
to be a measure of the energy available to do work.
Data on the left side of this figure
correspond to relatively small
values of Q. They therefore
describe systems in which there is
far more reactant than product.
The sign of ΔG for these systems is
negative and the magnitude of Δ G
is large. The system is therefore
relatively far from equilibrium and
the reaction must shift to the right
Q=partial P of
to reach equilibrium.
Pr/ partial P of
Re
Data on the far right side of this
figure describe systems in which
there is more product than
reactant. The sign of ΔG is now
positive and the magnitude of Δ G
is moderately large. The sign of Δ G
tells us that the reaction would
have to shift to the left to reach
equilibrium. The magnitude of Δ G
tells us that we don't have quite as
Q=partial P of
far to go to reach equilibrium.
Pr/ partial P of
Re
The points at which the straight line
in the above figure cross the
horizontal and vertical axes of this
diagram are particularly important.
The straight line crosses the vertical
axis when the reaction quotient for
the system is equal to 1 (as the
example is gas, so all at 1
atmosphere at standard condition
and 25°C, we will consider partial
pressure). This point therefore
describes the standard-state
conditions, and the value of ΔG at
When QP= 1: Δ G= this point is equal to the
ΔGo standard-state free energy of
reaction, Δ Go.
The point at which the straight line
crosses the horizontal axis
describes a system for which ΔG is
equal to zero. Because there is no
driving force behind the reaction,
the system must be at equilibrium.
When Qp = Kp: ΔG = 0
• http://chemed.chem.purdue.edu/genchem/to
picreview/bp/ch21/entropy.php
Some definitions
In a closed system, there cannot be a loss or gain of mass, but there can be a
change in energy, dE.
dE = dQ - dW (1)
Q=heat
W=work done by the system
W=force x distance.
Pressure, P= Force/surface area,
Force = P x surface area,
Equation (6) tells us that phases with small volume are favored at
higher pressure, and equation
(7) tells us that phases with high entropy (high disorder) are favored
at higher temperature.
Equation (5), dG = VdP – SdT,
tells us that the Gibbs Free
Energy is a function of P and T.
GA= GB at PE
. At pressures greater than P
phase B is
stable because it has a lower G than
phase A. At pressures less than PE
phase A is stable because it has a
lower G than phase B.
Finally, we look at a cross-section
across the bottom of the first
figure. Here we project the line of
intersection of the Free Energy
surfaces onto the P - T plane. Along
the line of intersection of the
surfaces GA= GB
. The line separates two fields, one
at low P in
which A is the stable phase and one
at higher P in which phase B
B
is stable. This is a classic P-T phase
diagram. The line represents
all values of P and T where
For a reaction, A ↔B
dΔG = ΔVdP - ΔSdT (8)
ΔG = the change in Gibbs Free Energy of the reaction
ΔS = the change in entropy of the reaction
and ΔV = the change in volume of the reaction
( 9)
This relation is known as the Clausius - Clapeyron
Equation. It is important because it tells us the
slope of the equilibrium boundary or reaction
boundary on a Pressure versus Temperature
phase diagram.
ΔS is positive.-RIGHT???
ΔV is also positive-?????
is also positive.
Note that the reaction boundary for
Andalusite <=> Sillimanite
has a negative slope.
ΔV is negative or postive?
ΔS s positive or negative?
dP/dT negative.
Devolatization Reactions
ΔG = ∑ GPr- ∑ Gre
ΔS=∑ SPr- ∑ Sre
ΔH= ∑ HPr- ∑ Hre
In general ΔG, ΔH, ΔS, and ΔV are dependent of Pressure and Temperature, but at any
given T & P:
If ΔG < 0 (negative) the chemical reaction will be spontaneous and run to the right,
If ΔG = 0 the reactants are in equilibrium with products,
and if ΔG > 0 (positive) the reaction will run from right to left.
● G is a measure of relative chemical stability for a
phase
● We can determine G for any phase by measuring H
and S for the reaction creating the phase from the
elements
● We can then determine G at any T and P
mathematically
● Most accurate if we know how V and S vary
with P and T
• dV/dP is the coefficient of isothermal
compressibility
• dS/dT is the heat capacity (Cp)
Temperature Dependence of G, H, and S
As stated above, G, H, and S depend on Temperature and Pressure. But,
because G depends on
H and S, it is usually more convenient to consider the temperature dependence
of H and S, so
that if we know H and S at any given temperature, we can calculate G.
where Cp is the heat capacity at constant pressure. The heat capacity is the
amount of heat necessary to raise the temperature of the substance by 1 oK.
CP can be expressed in terms of power series f T
CP=a+bT+c/T2............
(a+bT+c/T2) dT
[(a+bT+c/T2) /T] dT
G P −G P =
2 1 z
P2
P1
VdP
Ideal Gas
– As P increases V decreases
– PV=nRT Ideal Gas Law
• P = pressure
• V = volume
• T = temperature
• n = # of moles of gas
• R = gas constant
= 8.3144 J mol-1 K-1
Figure 5.5. Piston-and-cylinder apparatus to
compress a gas. Winter (2010) An Introduction to
Igneous and Metamorphic Petrology. Prentice Hall.
P x V is a constant at constant T
Gas Pressure-Volume Relationships
Since G P −G P =
2 1 z
P2
P1
VdP
G P −G P =
2 1 z
P2
P1
RT
P
dP
and, since R and T are certainly independent of P:
G P − G P = RT
2 1 z P2
P1
1
P
dP
Gas Pressure-Volume Relationships
And since z 1
x
dx = ln x
G of a gas at some P and T = G in the reference state (same T and 0.1 MPa)
+ a pressure term
Gas Pressure-Volume Relationships
The form of this equation is very useful
GOl=nFoµFo+nFaµFa
If ‘N’=∑n=ni+nj+nk+.....
Gα/N=∑(ni/N)µi
=
∑Xiµi (X=Mole fracion)
Gα/N= Molar Gibb’s free energy
Intensive or extensive variable?
For pure phase Xi=1 as (ni/ni),
Therefore, Molar Gibb’s free energy=µi for
pure phase
Fundamental Equation of Chemical Thermodynamics
Gibbs free energy will be of most interest to us, since P and T are the most obvious
choices of independent variables for geologic application
Open System
Gibbs free energy will be of most interest to us, since P and T are the most obvious
choices of independent variables for geologic application
Close System
Diffusion
• At equilibrium- µi of component i must be
same in every coexisting phases that contain
it.
• If µi is lower in one phase that the other, free
energy of system could be lowered by
migration of the component from a phase
with higher µi to a phase with lower µi
• Difference in µi drives diffusion.
Response of system when A
or B atom is introduced-
Calculate the XB and draw a
tangent from the point on
the curve that corresponds
to the specific XB.
The µA and µB are the
corresponding chemical
potential of the A and B
phases.,
P, T constant
If homogenized mixture has lower energy (X0)
Location 2-
µ A< µ B
Location 1-
µ A> µ B
And
P, T constant
Till the system become homogenous X0 to minimize
energy.
If homogenized mixture has higher energy (X0)
Location 2-
µA>µB
Location 1-
µ A< µ B
And
A will flow from 2 to 1
Figure 27.3. Activity-composition relationships for the enstatite-ferrosilite mixture in orthopyroxene at 600 oC and 800oC. Circles are data
from Saxena and Ghose (1971); curves are model for sites as simple mixtures (from Saxena, 1973) Thermodynamics of Rock-Forming
Crystalline Solutions. Winter (2010) An Introduction to Igneous and Metamorphic Petrology. Prentice Hall.
Why activity or fugacity?
Why activity or fugacity?
There are two ways to deal with real systems that deviate appreciably from
ideal conditions:
1.Use a more accurate phenomenologically (i.e., real) Equation of State like the
van der Waal' that models the system more accurately. This must explicitly
address intermolecular forces and other effects that exist in real system and the
concentration used is 1 mol/L (i.e., the real concentration).
2.Use an ideal equation pf state like the ideal gas Equation of State in, but use an
"effective concentration" of 0.816 mol/L to generate the observed pressure (that
is, the gas behaves as if it has a reduced concentration of 100%-18.36% =
81.6% of the real concentration).
Why activity or fugacity?
The activity of a substance describes the effective concentration
of that substance in the reaction mixture. Activity takes into
account the non-ideality of the reaction mixture, including
solvent-solvent, solvent-solute, and solute-solute interactions.
Thus, activity provides a more accurate description of how all of
the particles act in solution.
Figure 27.3. Activity-composition relationships for the enstatite-ferrosilite mixture in orthopyroxene at 600 oC and 800oC. Circles are data
from Saxena and Ghose (1971); curves are model for sites as simple mixtures (from Saxena, 1973) Thermodynamics of Rock-Forming
Crystalline Solutions. Winter (2010) An Introduction to Igneous and Metamorphic Petrology. Prentice Hall.
Activity Models (Activity-Composition
Relations) for Crystalline Solutions
• aprp = γMg3XMg3
• aalm = γ Fe3XFe3
• agrs = γ Ca3XCa3
• asps = γ Mn3XMn3
If we assume ideal behavior (γ = 1) in garnet
• aalm = Xalm3 = [Fe/(Fe + Mg + Ca + Mn)]3
• aprp = Xprp3 = [Mg/(Fe + Mg+ Ca + Mn)]3
......................................................
The Equilibrium Constant
Figure 27.4. P-T phase diagram for the reaction Jadeite + Quartz = Albite for various values of K. The equilibrium curve for K = 1.0 is the
reaction for pure end-member minerals (Figure 27.1). Data from SUPCRT (Helgeson et al., 1978). Winter (2010) An Introduction to
Igneous and Metamorphic Petrology. Prentice Hall.
Geothermobarometry
Use measured distribution of elements in coexisting
phases from experiments at known P and T to estimate
P and T of equilibrium in natural samples
Thermometry
Exchange Reactions
• Many thermometers are based on exchange
reactions, which are reactions that exchange
elements but preserve reactant and product
phases.
• For example: Fe3Al2Si3O12+ KMg3AlSi3O10(OH)2 =
Mg3Al2Si3O12 + KFe3AlSi3O10(OH)2
• almandine + phlogopite = pyrope + annite
We can reduce this reaction to a simple exchange
vector: (FeMg)gar+1 = (FeMg)bio-1
• Popular thermometers include garnet-biotite (GARB),
garnet-clinopyroxene, garnet-hornblende, and
clinopyroxene-orthopyroxene; all of these are based
on the exchange of Fe and Mg, and are excellent
thermometers because ΔrV is small, such that
dP/dT= ΔrS/ ΔrV is large (i.e., the reactions have steep
slopes and are little influenced by pressure).
Let's write the equilibrium constant for this exchange
reaction K = (aprpaann)/(aalmaphl)
• thus ΔrG = -RT ln (aprpaann)/(aalmaphl)
• If we assume ideal behaviour, The equilibrium constant is
K = KD=(XMggar XFebio)/(XFegar XMgbio)
• When discussing element partitioning it is common to
define a distribution coefficient KD, which is just the
equilibrium constant without the exponent (this just
describes the partitioning of elements and not the
partitioning of chemical potential):
KD = (XMggar XFebio)/(XFegar XMgbio) = (Mg/Fe)gar /(Mg/Fe)bio =
K1/3
• John Ferry and Frank Spear in 1978 measured experimentally the
distribution of Fe and Mg between biotite and garnet at 2 kbar and
found the following relationship:
Geothermobarometry
The Garnet - Biotite geothermometer
K=KD.Kγ, therefore, lnK= lnKD + lnKγ
• If you compare their empirical equation
If we compare this relation with
ln KD = -2109 / T + 0.782 ln K = - ( G° / RT) = -( H / RT) - (P V /
r r r
RT) + (rS / R)
As in case of ideal solution, ln K =ln
KD
or rH
As value of m (tanangent of its inclination, m = tan θ) decreases angle decreases wrt to
x axis (which is T in a P-T diagram)
Therefore, if the reaction has lower value for rV, the slope [dP/dT=∆S/ ∆V(rV)] will be
high, so the angle with T will be increasing- very sensitive to temperature, good
thermometry
Therefore, if the reaction has larger value for rV, the slope will be small, so the angle
with T will be decreasing- insensitive to temperature good barometry
Volume change
Figure 27.5. Graph of lnK vs. 1/T (in Kelvins) for the Ferry and Spear (1978) garnet-biotite exchange equilibrium at 0.2 GPa from Table
27.2. Winter (2010) An Introduction to Igneous and Metamorphic Petrology. Prentice Hall.
• To plot the KD lines in PT space
•
Geothermobarometry
The Garnet - Biotite geothermometer
Figure 27.6. AFM projections showing the relative distribution of Fe and Mg in garnet vs. biotite at approximately 500 oC (a) and 800oC (b).
From Spear (1993) Metamorphic Phase Equilibria and Pressure-Temperature-Time Paths. Mineral. Soc. Amer. Monograph 1.
Barometry
• Net-Transfer Reactions
Net-transfer reactions are those that cause phases to
appear or disappear. Geobarometers are often based on net-
transfer reactions because ΔrV is large and relatively insensitive to
temperature.
A popular one is GASP:
3CaAl2Si2O8 = Ca3Al2Si3O12 + 2Al2SiO5 + SiO2
anorthite = grossular + kyanite + quartz
Figure 27.14. Chemically zoned plagioclase and poikiloblastic garnet from meta-pelitic sample 3, Wopmay Orogen, Canada. a. Chemical
profiles across a garnet (rim → rim). b. An-content of plagioclase inclusions in garnet and corresponding zonation in neighboring plagioclase.
After St-Onge (1987) J. Petrol. 28, 1-22 .
Geothermobarometry
P-T-t Paths
Figure 27.15. The results of applying the garnet-biotite geothermometer of Hodges and Spear (1982) and the GASP geobarometer of Koziol
(1988, in Spear 1993) to the core, interior, and rim composition data of St-Onge (1987). The three intersection points yield P-T estimates
which define a P-T-t path for the growing minerals showing near-isothermal decompression. After Spear (1993).
What is a P-T-t path
Metamorphism is a
• Dsc xx dynamic process,
involving changes in
temperature ± pressure
through time. The
pressure (P) -
temperature (T) - time (t)
A common pressure-temperature path for regional metamorphism. The
rate of prograde metamorphism (heating) and rate of retrograde
path of a metamorphic
metamorphism (cooling) may not be the same. The duration of the path
from start (onset of metamorphism) to finish (exposure of the rock at the
rock is the set of all P-T
Earth's surface) will vary from rock to rock depending on the tectonic
history
conditions experienced
by a rock during its
metamorphic history
What are some common P-T paths?
• Some rocks may record more of their P-T paths. If a rock contains a partial record of
its P-T path, this is both good and bad. This is good because we want as much P-T
path information as possible, so as to be able to interpret the thermal/tectonic
processes and history as well as possible. This is bad because the mineralogical and
textural evidence for P-T path segments other than the conditions of the peak of
metamorphism represent disequilibrium. In most cases, however, the evidence for
disequilibrium can be very useful because it can be used to reconstruct P-T path
segments.
• A few common methods for inferring P-T path
segments are:
1. Mineral inclusions
Some minerals contain inclusions of other minerals.
For example, garnets commonly contain inclusions of
minerals that were present in the rock matrix as the
garnet grew, but that were not completely eliminated
by metamorphic reactions during progressive
metamorphism. The growing garnets surrounded
these relict minerals as the garnets grew, and the
relict minerals are preserved as mineralogical evidence
of an earlier stage in the metamorphic history of the
rock.
Figure 3. Left: Photomicrograph (plane light) of kyanite inclusions in garnet; field of view = 2 mm. Right: Photomicrograph (crossed polars) of
sillimanite in the matrix; field of view = 4 mm.
Figure 27.14. Chemically zoned plagioclase and poikiloblastic garnet from meta-pelitic sample 3, Wopmay Orogen, Canada. a. Chemical
profiles across a garnet (rim → rim). b. An-content of plagioclase inclusions in garnet and corresponding zonation in neighboring plagioclase.
After St-Onge (1987) J. Petrol. 28, 1-22 .
Figure 5. Common zoned minerals. (a) False color X-ray map showing Mn distribution
in a garnet from Iran (Sepahi et al., 2003). The garnet core contains more Mn than
the garnet rim (or the matrix); this is typical growth zoning. The garnet is 1.3 mm in
diameter; (b) Photomicrograph (crossed polarized light) showing zoning in
plagioclase in a metamorphosed igneous rock; oscillatory zoning is common in
igneous plagioclase, but metamorphic plagioclase may also be zoned in the anorthite
(Ca) and albite (Na) components; field of view = 2 mm; (c) Photomicrograph (plane
polarized light) of a zoned tourmaline crystal in a kyanite schist; field of view = 2 mm;
(d) Cathodoluminescence image of isotopically zoned zircon from the Nigde Massif,
Turkey, with the U-Pb age of the core and rim labeled in millions of years (Whitney
et al., 2003).
• Using zoning information to reconstruct the part of the P-T path
experienced by the zoned mineral is not simple, but a few general aspects
of the relationship of zoning to P-T path may easily be inferred:
(a) Garnets with Mn-rich cores and Mn-poorer rims record growth zoning
that represents the change from the lower-T conditions at which the
garnet core grew to the higher-T conditions at which the garnet rim grew
(i.e., prograde metamorphism involving increasing temperature and
pressure). Mn is preferentially partitioned into garnet relative to most
other common minerals, so Mn is sequestered in early-formed garnet,
depleting the local environment of the growing garnet in Mn.
(b) Minerals that show major element growth zoning probably did not
experience very high metamorphic temperatures. At high temperature (>
700 C) and sufficient duration, zoning may be homogenized as
intracrystalline diffusion becomes more effective at eliminating
compositional variation. An unzoned mineral that is typically zoned at
low-medium metamorphic grades has either experienced high
temperature conditions or was never zoned (owing to a simple reaction
history at limited P-T or to growth entirely at high-T).
3. Reaction textures
• Some metamorphic rocks contain evidence for incomplete
reactions or other textural evidence for part of the P-T path. A
simple example is the partial replacement of andalusite by
sillimanite (Fig. 6a) by the polymorphic transformation of
andalusite to sillimanite. Textural evidence may also be useful in
cases where the reactants have been completely consumed if the
shape of one or more reactants are preserved; for example,
pseudomorphs(Figs. 6b-d).
An example of a more complex reaction texture involves the
formation of coronas , which consist of one or more shells (rims,
moats) of a mineral or minerals around a central (reactant) phase
(Fig. 6e). In many cases, coronas also involve the fine-scale
intergrowth of minerals in a texture known as symplectite(Figs.
6e-f).
Figure 6. Images of reaction textures. (a) Photomicrograph (plane light) showing the partial replacement of andalusite by
sillimanite in a schist from Iran. Note: The crystallization sequence (sillimanite after andalusite) can't be inferred only from this
photo; (b) Photomicrograph (plane light) showing the complete replacement of kyanite by sillimanite in a sample of gneiss from
the Thor-Odin dome, British Columbia. The former presence of kyanite is known because some pseudomorphs (not shown)
contain relict kyanite. Without these relics, and based only on the tabular shape of the pseudomorph, it would be difficult to
infer whether the original mineral was kyanite or andalusite; (c) Crossed polar view of (b) showing the randomly oriented
sillimanite in the pseudomorph; (d) Photomicrograph (plane light) showing a partial pseudomorph of chlorite after garnet from a
retrograded eclogite, Turkey. Garnet relics are present, but the former presence of garnet is clear also from the shape of the
pseudomorph; (e) Photomicrograph (plane light) showing a corona texture from a Thor-Odin dome gneiss. The central Al2SiO5
phase (sillimanite after kyanite) is rimmed by an inner shell of spinel + cordierite symplectite and an outer shell of cordierit; (f)
Backscattered electron image of spinel (brightest phase) + cordierite (darkest gray) + anorthite (medium gray) from a Thor-Odin
symplectite.
How are P-T-t paths interpreted?
• Perhaps the best way to understand entropy as a driving force in nature is to conduct a simple
experiment with a new deck of cards. Open the deck, remove the jokers, and then turn the deck
so that you can read the cards. The top card will be the ace of spades, followed by the two, three,
and four of spades, and so on. Now divide the cards in half, shuffle the deck, and note that the
deck becomes more disordered. The more often the deck is shuffled, the more disordered it
becomes.What makes a deck of cards become more disordered when shuffled?
• In 1877 Ludwig Boltzmann provided a basis for answering this question when he introduced the
concept of the entropy of a system as a measure of the amount of disorder in the system. A
deck of cards fresh from the manufacturer is perfectly ordered and the entropy of this system is
zero. When the deck is shuffled, the entropy of the system increases as the deck becomes more
disordered.
• There are 8.066 x 1067 different ways of organizing a
deck of cards. The probability of obtaining any
particular sequence of cards when the deck is
shuffled is therefore 1 part in 8.066 x 1067. In theory, it is
possible to shuffle a deck of cards until the cards fall into perfect order. But it isn't very likely!
• Boltzmann proposed the following equation to describe the relationship between entropy and the
amount of disorder in a system.
• S = k ln W
• In this equation, S is the entropy of the system, k is a proportionality constant equal to the ideal
gas constant divided by Avogadro's constant, ln represents a logarithm to the base e, and W is the
number of equivalent ways of describing the state of the system. According to this equation, the
entropy of a system increases as the number of equivalent ways of describing the state of the
system increases.
• The relationship between the number of equivalent ways of describing a system and the amount
of disorder in the system can be demonstrated with another analogy based on a deck of cards.
There are 2,598,960 different hands that could be dealt in a game of five-card poker. More than
half of these hands are essentially worthless. Winning hands are much rarer. Only 3,744
combinations correspond to a "full house," for example. The table below gives the number of
equivalent combinations of cards for each category of poker hand, which is the value of W for this
category. As the hand becomes more disordered, the value of W increases, and the hand becomes
intrinsically less valuable.
Number of Equivalent Combinations for Various Types of Poker Hands
Hand W ln W
Royal flush
(AKQJ10 in one 4 1.39
suit)
Straight flush (five
cards in sequence 36 3.58
in one suit)
Four of a kind 624 6.44
Full house (three
of a kind plus a 3,744 8.23
pair)
Flush (five cards in
5,108 8.54
the same suit)
Straight (five cards
10,200 9.23
in sequence)
Three of a kind 54,912 10.91
Two pairs 123,552 11.72
One pair 1,098,240 13.91
No pairs 1,302,540 14.08
The Relationship Between Free Energy
and Cell Potential
• The value of ΔG for a reaction at any moment in time
tells us two things. The sign of ΔG tells us in what
direction the reaction has to shift to reach equilibrium.
The magnitude of ΔG tells us how far the reaction is
from equilibrium at that moment.
• We can now compare it with the equation used to describe the relationship between
the free energy of reaction at any moment in time and the standard-state free
energy of reaction.
ΔG = ΔGo + RT ln Q
• These equations are similar because the Nernst equation is a special case of the
more general free energy relationship. We can convert one of these equations to the
other by taking advantage of the following relationships between the free energy of
a reaction and the cell potential of the reaction when it is run as an electrochemical
cell.
ΔG = -nFE
• Therefore, we introduce the concept of
enthalpy (H), which is the sum of the internal
energy of the system plus the product of the
pressure of the gas in the system times the
volume of the system.
H = E + PV
Δ H = Δ E + Δ (P V)
The Relationship Between Free Energy
and Cell Potential
• The value of ΔG for a reaction at any moment in time
tells us two things. The sign of ΔG tells us in what
direction the reaction has to shift to reach equilibrium.
The magnitude of ΔG tells us how far the reaction is
from equilibrium at that moment.
• We can now compare it with the equation used to describe the relationship between
the free energy of reaction at any moment in time and the standard-state free
energy of reaction.
ΔG = ΔGo + RT ln Q
• These equations are similar because the Nernst equation is a special case of the
more general free energy relationship. We can convert one of these equations to the
other by taking advantage of the following relationships between the free energy of
a reaction and the cell potential of the reaction when it is run as an electrochemical
cell.
ΔG = -nFE
• A good example of this phenomenon is the reaction in which NO2
dimerizes to form N2O4.
There have been many attempts to build a device that violates the
laws of thermodynamics. All have failed. Thermodynamics is one of the
few areas of science in which there are no exceptions.
• For the sake of argument, let's assume that there is no significant
change in either ΔHo or ΔSo as the system is cooled. The
contribution to the free energy of the reaction from the enthalpy
term is therefore constant, but the contribution from the entropy
term becomes smaller as the temperature is lowered.