2 c125 Module1 w25 Final
2 c125 Module1 w25 Final
In this module, we will learn introductory concepts of chemical thermodynamics, the branch of physical
chemistry that deals with the study of energy changes physical processes and chemical reactions.
Thermodynamics helps us understand how energy flows and is transformed in chemical systems, and how it
influences the behavior and spontaneity of reactions.
Let’s consider a system which undergoes a transformation between two states A and B. The transformation may
be physical or chemical, and may occur in one steps or many steps.
In a physical change, there is no change in the chemical composition of the system. However, there may be
changes in temperature, pressure, volume, density, or the phase (state of aggregation). Examples of physical
transformations include heating, cooling, expansion, compression, and phase changes, such as
vaporization, fusion (melting), and sublimation. In a chemical change, the chemical composition of the
system is changed by a chemical reaction.
The concepts of energy and entropy are central to understanding and predicting physical and chemical
changes. For all physical and chemical changes:
• Mass is conserved.
• Energy is conserved. (This is the first law of thermodynamics.)
• Entropy does not decrease. (This is the second law of thermodynamics.)
By the end of this module, you should be able to describe the energy and entropy changes that accompany
physical and chemical changes and predict the direction of spontaneous change.
CHEM 123/125 Module 1: Energy, Entropy, and Spontaneous Change page 2 of 42
Before we can state the laws of thermodynamics in a form that is useful for chemistry, we
must review some basic ideas, and discuss in detail the concepts of heat and work.
surroundings − a region outside of the system that serves as a place to make measurements
of the system
Types of systems
A system may be classified as open, closed or isolated depending on whether matter or energy is permitted to
cross the boundaries and be exchanged with the surroundings.
Heat, q
Heat is energy in transition. Heat, q, is energy that flows from a region of high temperature to a region of low
temperature. Heat will flow between a system and its surroundings when the system is not in thermal
equilibrium with its surroundings (or is not an isolated system). When the system and surroundings have the
same temperature, they are said to be in thermal equilibrium with each other.
q>0 When the value of q is positive: heat is absorbed by the system from the surroundings and this
process is defined as being endothermic (e.g. the melting ice or the vaporization of water are both
endothermic processes)
q<0 When the value of q is negative: heat is released by the system into the surroundings and this process
is defined as being exothermic (e.g. combustion reactions are exothermic)
In a hot object, the molecules have, on average, high kinetic energies. When a hot object is placed in contact
with a cold object, energy is transferred from high energy molecules to low energy molecules via molecular
collisions.
• temperature increases
The temperature increase may or may not be accompanied by a volume change. When the temperature
increases, the average kinetic energy of the molecules also increases. So, when heat causes a temperature
increase, most of the heat is used to increase the kinetic energies of the molecules.
A phase change occurs at a constant temperature (e.g. water boils at 100 oC). The heat is used to
overcome the attractive forces that cause molecules to aggregate. So, when heat causes a phase
change, the heat is used to increase the potential energies of the molecules.
diathermic − a boundary that permits heat to be transferred between the system and its surroundings (e.g. a
stainless steel container has diathermic walls)
adiabatic − a boundary that does not permit heat to be transferred between the system and its surroundings (e.g. a
thermos with evacuated walls used to keep soup or coffee hot for an extended period). If a process is
adiabatic, no heat is transferred and q = 0.
CHEM 123/125 Module 1: Energy, Entropy, and Spontaneous Change page 4 of 42
Imagine that we have a solid that we heat continuously at constant pressure from an initial temperature Ti to a
final temperature Tf. The temperature of the substance will change as shown in the diagram below.
Tf
warming
vapour ∆Tgas
Temperature
warming
liquid ∆Tliquid
Heat added
• The temperature of the system remains constant while the system is undergoing a phase change.
• The heat of vaporization (qvap) is typically much greater than the heat of fusion (qfus). Why?
What information about the substance can we obtain from heating curve?
1. The heat capacities of solid, liquid and gas can be obtained from the slopes of the ◄ Heat capacity provides a
measure of how much heat
“warming lines”. (We’ll say more about heat capacity in the next section.) is required to increase the
temperature by one degree.
2. The heat of fusion (qfus) and the heat of vaporization (qvap) can be
obtained from the horizontal portions of the heating curve.
qsolid
Csolid = heat capacity of solid = qsolid = heat absorbed by solid ∆Tsolid = Tfus − Ti
Tsolid
qliquid
Cliquid = heat capacity of liquid = qliquid = heat absorbed by liquid ∆Tliquid = Tvap − Tfus
Tliquid
qgas
Cgas = heat capacity of gas = qgas = heat absorbed by gas ∆Tgas = Tf − Tvap
Tgas
CHEM 123/125 Module 1: Energy, Entropy, and Spontaneous Change page 5 of 42
Heat Capacity
The heat capacity of a substance (or a system) is the amount of heat required Note: A temperature change in K is
equal to the temperature change in oC.
to raise the temperature by 1 oC or 1 K. We can define an “average” heat It is straightforward to demonstrate this.
ΔT / K = Δt / oC
To measure heat capacity, we deliver a known quantity of heat and measure the
temperature change produced. The heat capacity depends not only on the
“nature” of the sample or object, but also on the size of that sample or object.
Once we know the heat capacity, we can use it to convert a temperature change
into a quantity of heat, or vice versa.
The heat capacity of an object (e.g. a reaction vessel) is given in J K−1. The
heat capacity of a pure substance is typically given in J K−1 g−1
or J K−1 mol−1. For pure substances, we can write:
c = specific heat capacity
(in J K−1 g−1)
q = m c T = n C T (pure substances only)
C (or Cm) = molar heat capacity
(in J K−1 mol−1)
Heat capacity also depends on whether the heating occurs at constant volume,
or at constant pressure. Therefore, in principle, we should distinguish between The numerical value of the heat capacity
is the same whether we express the
heat capacities for constant volume heating and constant pressure heating. temperature in K or oC.
For example:
CV = heat capacity for heating at constant volume
C = 10 J K−1 mol-1 = 10 J oC−1 mol−1
Cp = heat capacity for heating at constant pressure
The reason is that, for heat capacity, K
and oC refer to a temperature change. As
Constant pressure molar heat capacities, near room temperature, are given in we mentioned above, a temperature
change is the same whether it is
the table on the right for a few different substances. However, the heat expressed in K or oC.
capacity of a substance does change with temperature.
Learning Tip: When a substance is heated, the added energy Substance C p / J K−1 mol−1
is absorbed by the atoms or molecules. In general, the greater at room temperature
the number of ways these entities can absorb energy, the N2(g) 29.1
higher the molar heat capacity. A substance containing O2(g) 29.4
molecules with complex structures has more ways to absorb H2O(l) 75.3
energy and therefore, a higher heat capacity than a substance CCl4(l) 130.7
made up of molecules with simpler structures. I2(s) 54.4
C(graphite) 8.5
Cu(s) 24.4
CHEM 123/125 Module 1: Energy, Entropy, and Spontaneous Change page 6 of 42
Example 1-1: What is the final temperature of the system when 0.500 g of H2O(l), initially at 20 °C, is mixed
with 0.750 g of Cu(s) initially at 85 °C? Assume all processes occur at constant pressure.
Data: 𝐶𝑝̅ = 24.4 J K−1 mol−1 for Cu(s) and 𝐶𝑝̅ = 75.3 J K−1 mol−1 for H2O(l).
Work, w
In physics, work is represented by the product of a force and a displacement. In chemical systems, the
displacement is associated with a change in the relative positions or configuration of the particles of the system.
Work, w, can take on many different forms. Here are some examples.
• expansion work, which can done by an expanding gas that pushes a piston to lift a mass against gravity
• electrical work, which can be done by moving a charged particle, such as an electron, from a region of high
electrical potential to a region of low electrical potential
• surface tension work, can be done when the surface area of a liquid changes.
We are interested exclusively in expansion work. It comes into play because our atmosphere exerts a continuous
downward pressure on any system. Therefore, when the pressure exerted by the surroundings causes a decrease
in the volume of the system, then the surroundings have done work on the system. When the volume of a system
changes by an amount V = Vf − Vi in the presence of a constant external pressure, Pex, then the work done by the
surroundings on the system is defined to be:
◄ This equation for w only applies if the external pressure does not
Expansion work: w = −Pex V change throughout the volume change. In the event that the
external pressure changes, we need to use a more general
expression which is based on dividing the total volume change into
an infinite number of infinitesimally small volume changes, dV.
Convince yourself that the product of pressure and volume gives you units of
Pressure units:
energy. (The conversion factors on the right should help.)
1 atm = 101.325 kPa (exactly)
The minus sign in the definition for w arises from the fact that the pressure
1 bar = 100 kPa
exerted by the surroundings opposes an increase in volume (i.e. the force
Volume units:
vector opposes the displacement vector). Consequently, the surroundings do
1 L = 10−3 m3
positive work when the system is compressed and negative work when the
system expands. Energy units:
1 J = 1 N m = 1 kg m2 s−2
1 Pa m3 = 1 kPa L = 1 J
Sign conventions for work, w:
1 bar L = 100 J
w>0 When the value of w is positive work is done on the 1 atm L = 101.325 J
system by the surroundings (e.g. the compression of gas 1 cal = 4.184 J (exactly)
inside a container)
▲ Did you know? 1 “food” calorie is
actually 1 kilocalorie.
w<0 When the value of w is negative work is done on the
surroundings by the system (e.g. the expansion of gas
generated during an explosion)
CHEM 123/125 Module 1: Energy, Entropy, and Spontaneous Change page 8 of 42
U = q + w
Remarks:
1. The First Law provides no definition or interpretation of U. It provides only a “recipe” for calculating
changes in U. For chemical systems, interal energy, U refers to the total energy contained within the
system, which includes all forms of energy at the microscopic level. We will soon learn that this energy
is associated with the motion and interactions of theparticles that make up the system. Internal energy
includes both the kinetic energy (from the motion of the particles) and the potential energy (from the
interactions between the particles).
2. According to the First Law, a given internal energy change (U) can be achieved via heat transfer,
doing work, or through some combination of the two. However, heat and work affect the molecular
motions in different ways. When energy is transferred as work, molecular motions tend to change in an
organized fashion. When energy is transferred as heat, the molecular motions are affected in random
fashion.
Example 1-3: (a) What are q, w, and U if 125 kJ of work is done on a system and 275 kJ of heat flows into
the surroundings? (b) How much work must be done to achieve the same internal energy change under
adiabatic conditions?
CHEM 123/125 Module 1: Energy, Entropy, and Spontaneous Change page 10 of 42
Chemical processes are commonly carried out at constant volume (e.g. in a rigid, sealed vessel), or at
constant atmospheric pressure (e.g. “open” to the atmosphere). We can derive an alternative statements of
the First Law for these two important situations.
Key Concept: When a reaction occurs at constant volume, the heat transferred is equal to the internal energy
change for the system, qv =U.
Key Concept: When a reaction occurs at constant pressure, the heat transferred is equal to the enthalpy
change for the system, qp =H.
Note: In a constant pressure reaction, the volume of the system may change.
= qp − Pf Vf + Pi Vi ◄ Here, we have used the fact that Pex = Pi and Pex = Pf.
Uf + Pf Vf = qp + Ui + Pi Vi
Let’s define the enthalpy (H) of a system as H = U + PV. Then: ◄ The enthalpy, H, has no physical
interpretation. For a constant pressure
Hf = qp + Hi
process, the enthalpy change does have
H = qp a physical interpretation: ΔH represents
the heat transferred in a constant
pressure process.
CHEM 123/125 Module 1: Energy, Entropy, and Spontaneous Change page 12 of 42
The First Law helps us “keep track” of energy when a change occurs within a system of interest.
However, the First Law only requires that energy be conserved, it places no other restriction on
the energy transfer. The following processes do not violate the First Law (energy
conservation), yet they are never likely to be observed.
• A ball at rest on the floor absorbs energy from the floor and begins to bounce
higher and higher.
• A piece of hot copper in contact with a piece of cold zinc absorbs energy from ◄ FYI: The specific heat
capacities of copper and zinc
the zinc and becomes “hotter”, leaving the zinc even “colder”. are approximately equal.
ice at −5 oC
10 oC
25 oC
rust O2(g)
(Fe2O3)
Fe(s) Fe(s)
Predicting the direction of spontaneous change involves a system property called entropy. The entropy of a
system provides a direct measure of the number of ways a system can be “microscopically” different.
CHEM 123/125 Module 1: Energy, Entropy, and Spontaneous Change page 13 of 42
What is Entropy? Entropy is a measure of the number of possible microscopic states of a system. It
connects macroscopic thermodynamic properties (like temperature and pressure) with microscopic
behaviors (like the motion of molecules).
Consider, for example, a fixed amount of an ideal gas, as shown in the image below. The gas has a single
macroscopic state, characterized by the values of n, T, V and P. The state of the gas will not change without
some external influence (e.g. by adding more gas, increasing the temperature, compressing the gas, etc.)
However, on the microscopic level, the state of the system is not so easily characterized: the molecules are in
continuous random motion, experiencing collisions with each other or the walls of the container. The positions,
velocities and energies of individual molecules change from one instant to the next. Thus, for a given
macroscopic state, characterized by n, T, P, and V, there are many possible microscopic configurations
(microstates), each of which might be characterized by giving the position, velocity and energy of every
molecule in the gas. However, such a description is not consistent with quantum mechanics because, according
to the Heisenberg uncertainty principle, exact values for the position and velocity (or momentum) of a particle
cannot be simultaneously specified.
The Boltzmann equation for entropy provides the relationship between the number of microstates and entropy.
In the equation below, S is entropy, kB is the Boltzmann constant, and W is the number of microstates. The
Boltzmann constant is related to the gas constant R and Avogadro’s number, NA by the expression kB = R / NA.
We can think of kB as the gas constant per molecule.
The Boltzmann Equation for Entropy: S = kB ln W
Key Concept: The greater the number of microstates (W), the higher the entropy (S).
A microstate can be explained as a specific microscopic configuration describing how the particles of a
system are distributed among available energy levels.
Consider a system containing 3 distinguishable particles. The total energy of the system is 5E. How many
microstates are there?
Consider a system containing one mole of molecules, in the gas phase, at 298.15 K and 1 bar. A given
molecule carries a certain amount of energy and its energy may be the same as, but more likely different from,
that of another molecule. Furthermore, the energy of a given molecule will be distributed between various
types of motions. The translational, rotational, and vibrational motions are illustrated below for the CO
molecule.
What emerges from this view of molecules as energy carriers is that the number of microstates in a system
(and therefore, the entropy) increases with:
There are several equivalent ways of stating the second law. Here is the one that we’ll use.
ΔSuniv is the entropy change for the universe, ΔS is the entropy change for the system, and ΔSsurr is the
entropy change for the surroundings. This statement of the second law is not (yet) very helpful because it
presumes we know how to calculate entropy changes.
Remarks:
1. The statement given above for the Second Law is actually a corollary of (i.e. follows from) a more
fundamental statement of the Second Law. If you’re interested, see 13-2 in the 11th edition of Petrucci.
2. The Second Law provides no definition or interpretation of S. Our interpretation of the entropy comes from
our understanding of the molecular nature of matter. Entropy is a direct measure of the number of ways
a given system can be microscopically different because of the different ways of distributing the
system’s energy among the molecules of the system. It is interesting to note that the interpretation of
entropy came long after the discovery of S. (The Second Law of thermodynamics was discovered at a time
when there was considerable doubt about the reality of atoms.)
3. The second law provides a criterion for predicting the direction of change.
4. Thermodynamics provides formulas for calculating entropy changes for different physical and chemical
transformations. The derivations are very complicated. (If you’re interested, see section 13-2 in the 11th
edition of Petrucci. You are not responsible for the derivations.) We’ll simply accept the formulas below for
calculating entropy changes and use them to help us understand what types of changes cause an increase
in entropy and what types of changes cause a decrease in entropy.
CHEM 123/125 Module 1: Energy, Entropy, and Spontaneous Change page 17 of 42
The entropy change for the surroundings is determined by how much heat is absorbed or released by the
system.
q
Ssurr = −
Tsurr
An exothermic process releases heat into the surroundings (q < 0). The heat released into the
surroundings is taken up by (and is dispersed among) the molecules in the surroundings. Therefore,
an exothermic process increases the entropy of the surroundings ( Ssurr > 0).
Entropy changes for the system depend not only on the nature of system but also on the type of change.
tr H
Phase change S =
Ttr
Notes:
1. In the equation for S for a phase change, the symbol “tr” represents a transition. When applying
this equation, we specify which transition is involved by replacing “tr” with, for example, “fus” for
the melting of solid or “vap” for the vaporization of liquid. Also, this equation can be used only if
the phase transition occurs at the usual or normal transition temperature.
2. In the equations above, it is assumed that the heat capacity is constant (has the same value) for
all temperatures between Ti and Tf.
Increasing the temperature usually parallels an increase in the total energy available
to the molecules of the system. A greater number of energy levels is accessible to the
molecules as T increases. For a system containing a fixed number of molecules,
the number of microstates increases with the number of accessible energy
levels.
Let’s first consider the conversion of a liquid or solid to a gas at constant temperature and constant pressure.
In liquids and solids, the molecules have vibrational motion, and perhaps some rotational motion, but the
translational motion is somewhat restricted, especially in a solid. The molecules in a liquid or solid do not have
access to the huge number of very closely spaced energy levels (10 30 or more) that characterize the
translational motion of molecules in a gas. For vaporization and sublimation, a significant fraction of the
total energy becomes dispersed among a huge number of translational energy levels. Consequently, for
vaporization and sublimation, the entropy of the system increases.
The melting of a solid (fusion) also produces an increase in entropy, although the entropy change for fusion is
considerably smaller than it is for vaporization or sublimation. In a solid, molecules are restricted primarily to
vibrational motion. In a liquid, the molecules will also have some rotational and translational motion. When a
solid melts, the molecules gain some rotational and translational motion, so the energy of the system
becomes dispersed among a greater number of energy levels.
CHEM 123/125 Module 1: Energy, Entropy, and Spontaneous Change page 19 of 42
Example 1-5: (a) Without performing any calculations, predict whether ΔS is positive or negative when the
temperature of 1.00 mole of water is changed from 25 oC to 10 oC at a constant pressure of 1 atm.
(b) Calculate ΔS for this process.
CHEM 123/125 Module 1: Energy, Entropy, and Spontaneous Change page 20 of 42
Example 1-6: The conversion of 1.00 mole of H2O(l), initially at 25 oC and 1 atm, to
1.00 mol of H2O(g) at 115 oC and 1 atm involves three separate processes. Data for H2O(l):
C p = 75.29 J K−1 mol−1
H2O(l, 25 oC, 1 atm) ⎯⎯
→ H2O(l, 100 oC, 1 atm)
vapH = 40.7 kJ mol−1
H2O(l, 100 oC, 1 atm) ⎯⎯
→ H2O(g, 100 oC, 1 atm) (at 100 oC and 1 atm)
In chemistry, we often encounter processes that occur at constant temperature and pressure. For example,
phase changes and chemical reactions often occur at constant T and P. Chemists have a mechanism to
describe the Second Law of Thermodynamics in terms of the chemical system only, not the universe.
When temperature and pressure is constant, we can use an alternative statement of the Second Law, by first
defining something called Gibbs energy.
Gibbs energy: G H − TS
Gibbs Energy is a thermodynamic potential that helps predict the spontaneity of a process at constant
temperature and pressure by considering factors concerning only the system. The change in Gibbs Energy
(ΔG) during a reaction or process provides alternative criteria for describing spontaneity.
Criteria for spontaniety can be derived using the fundamental concepts that we have learned in this module.
Since we know the correlation between the sign of ΔSuniv and the spontaneous direction of a process, we can
derive the correlation between the sign of ΔG and the spontaneous direction.
𝛥G = 𝛥H − T𝛥S
= qp − T𝛥S
= − T𝛥Ssurr − T𝛥S
= − T(𝛥Ssurr + 𝛥S)
= − T𝛥Suniv
Since 𝜟𝑺𝒖𝒏𝒊𝒗 ≥ 𝟎, then: ΔG ≤ 0 for a process occurring at constant temperature and pressure.
CHEM 123/125 Module 1: Energy, Entropy, and Spontaneous Change page 22 of 42
So far, we have identified two related sets of criteria for identifying spontaneous change,
For each condition below, identify the sign of Suniv and G.
Suniv G
(general criteria) (Constant T and P)
It is important to note that when thermodynamic data is reported in tables, it is most commonly reported for the
positive value. You will be expected to apply a sign convention as necessary. For each of the phase changes
below, identify the expected sign of ∆H and ∆S.
CHEM 123/125 Module 1: Energy, Entropy, and Spontaneous Change page 23 of 42
Example 1-7: Calculate H, S, G, and Ssurr for the following process for
Data for H2O(l):
1.00 mol of H2O. Is the process thermodynamically reversible, irreversible, or C p = 75.3 J K−1 mol−1
impossible?
Data for H2O(s):
H2O(l, 263.15 K, 1 atm) ⎯⎯
→ H2O(s, 263.15 K, 1 atm) C p = 38.0 J K−1 mol−1
fusH = 6.01 kJ mol−1
(at 0 oC and 1 atm)
CHEM 123/125 Module 1: Energy, Entropy, and Spontaneous Change page 24 of 42
Ar(g) 154.84
Since the Third Law allows us to assign S = 0 at zero kelvin, we can C(gr.) 5.74
determine the value of S at any temperature. To understand why this is so, C(dia.) 2.38
imagine raising the temperature of a compound from 0 K to T while keeping CH4(g) 186.26
C6H6(l) 173.30
the pressure constant at P = 1 bar. If we assume that the compound is a
C6H6(g) 269.31
gas at the final temperature, then the following sequence of events occurs.
CO(g) 197.67
CO2(g) 213.74
1. Warm the solid from 0 K to its X(s) → X(s) Cu(s) 33.15
melting temperature, Tfus: 0K Tfus
Fe(s) 27.3
Fe2O3(s) 87.4
2. Melt the solid at its melting temperature: X(s) → X(l) H2(g) 130.57
Tfus Tfus H2O(l) 69.91
3. Warm the liquid from its melting temperature X(l) → X(l) H2O(g) 188.83
to its boiling temperature, Tvap: Tfus Tvap I2(s) 116.39
I2(g) 260.69
I(g) 180.79
4. Vaporize the liquid at its boiling temperature: X(l) → X(g)
Tvap Tvap N2(g) 191.50
NH3(g) 192.67
5. Warm the gas to the final temperature, T: X(g) → X(g) NO(g) 210.76
Tvap T
NO2(g) 240.06
O2(g) 205.04
The entropy changes for each of these processes can be calculated, so the
SO2(g) 248.22
total entropy change (per mole of substance) can also be calculated.
SO3(g) 256.76
S = S1o + S2o + S3o + S4o + S5o ▲ We will soon see that values can
be used to calculate entropy changes for
chemical reactions.
Since S = ST
o
− S0oK and S0oK = 0 J K −1 mol−1 , the value of S is equal to
o
the value of ST , the molar entropy at 1 bar and temperature T.
This approach is used to calculate S values at 298.15 K for all substances. Values are given
above for a few substances. Notice that the entropies of solids and liquids are typically much
smaller than the entropies of gases. When comparing substances in the same state of
aggregation, the entropy is greater for the substance having the more complicated structure.
CHEM 123/125 Module 1: Energy, Entropy, and Spontaneous Change page 25 of 42
▲ The standard molar entropy tends to increase with the structural complexity of the entities
making up the substance. The standard molar entropies of gases tend to be higher than those
o
of liquids, which in turn are higher than those of solids: Sgas Sliquid
o
Ssolid
o
.
Example 1-8 (On Your Own): Consider the following species. For ionic compounds, identify the ions
involved. For molecular compounds, draw proper Lewis structures. Based on the trends for standard molar
entropy, arrange the species in order of increasing standard molar entropy.
In all our previous examples, changes in U, H, S, and G arise from changes in temperature, pressure, or
volume or from a phase change. We now turn our attention to chemical transformations. In a chemical
transformation, changes in U, H, S, and G arise from changes in chemical composition although there may
also be changes in T, V, or p.
Reactants ⎯⎯⎯⎯
→ Products
T, V, p T , V , p
In this section, we focus on the calculation of rU, rH, rS, and rG for a reaction. We’ll consider:
• When the reaction conditions change, how do the values of rU, rH, rS, and rG change?
• What can we say about a system that has reached a state of chemical equilibrium?
For reactions, we focus on the quantities rH, rS, and rG rather than H, S, and G so we need to be clear
about the meaning of the “r” that appears as a subscript on the . Let’s focus on H and rH.
H represents the enthalpy change, in joules, of a system in which a chemical reaction occurs. The value of
H depends on the size of the system and is an extensive quantity.
rH is called the enthalpy of reaction. It represents the enthalpy change per mole of reaction, in joules per
mole. rH is the ratio of two extensive quantities (joules and moles). The ratio of two extensive quantities is an
intensive quantity. An intensive quantity does not depend on the size of the system.
▲ Density is another example of an intensive quantity. The density of water does not depend on the size of the system. For
example, the density of 1 mL of water is equal to the density of 1 L of water. Density is an intensive quantity because it is the ratio
of two extensive quantities (mass and volume).
1.8 Thermochemistry
If a chemical reaction takes place, the temperature of the system immediately after the
reaction is generally different from the temperature immediately before the reaction took
place. To restore the system to its original temperature, heat must flow from the system
into the surroundings or from the surroundings into the system.
endothermic reaction consumes heat; the heat required for the reaction
(qrxn > 0) flows into the system from the surroundings
In the laboratory, the majority of chemical reactions are carried out at constant
pressure (i.e. Pex = Pinitial = Pfinal = constant). Therefore, H = q. In general, the
enthalpy change depends on the nature of the reactants and products, and on the T
and P at which the reaction is carried out.
r H (T) is called the standard enthalpy of reaction. Unless specified otherwise, the
temperature is usually assumed to be 298.15 K and we write r H instead of r H (T ).
o
CHEM 123/125 Module 1: Energy, Entropy, and Spontaneous Change page 28 of 42
calorimeter = a device or reaction vessel that is thermally insulated from its surroundings
(no heat is lost to the surroundings)
There are many different types of calorimeters. Here are just a few. Schematic diagrams of a bomb
calorimeter and a coffee cup calorimeter are provided on the next page.
• bomb calorimeter good for studying combustion reactions, or other types of reactions
that release a lot of heat; a constant volume calorimeter
• coffee cup calorimeter quick and easy; for studying reactions in solution
In all cases, the calorimeter is thermally insulated from the surroundings. Therefore:
qrxn + qcal = 0 qrxn = − qcal
The heat absorbed or released by a reaction is deduced by measuring changes produced in the
calorimeter (e.g. “What temperature change is produced?”, or “How much ice melted?” or “How
much water froze?”)
When a reaction is carried out in a bomb calorimeter, the reaction occurs at constant volume. For a
constant volume process, we know that U = qV . Therefore, the experimentally measured heat of
However, if a reaction is carried out at constant pressure (for example, in a coffee cup calorimeter), then the
experimentally measured heat of reaction is the constant pressure heat of reaction, q p . Since H = q p for a
In general, the constant volume heat of reaction is not equal to the constant pressure heat of reaction. In other
words, qV q p . The difference between the two values is related to the amount of work that is done by (or on)
We can neglect the solids and liquids when calculating the quantity ∆(PV) = PfVf − PiVi. This is
because the volume of a liquid or solid is much smaller than that of a gas. If we focus on the gases
only, we have:
∆(PV) ≈ (PfVf)gases − (PiVi)gases
If we also assume that the gases behave ideally, (i.e., PV = nRT), then we can write:
In the equation above, ∆ngas is the change in the number of moles of gas for the reaction as written.
ngas = (c + d + ) − (a + b + )
the sum of coefficients the sum of coefficients
for the gaseous products for the gaseous reactants
When we substitute ∆(PV) = ∆ngas RT into the expression above for ∆H, we get the following equation.
H U + ngasRT
We established previously that ∆U = qv and ∆H = qp, so we can write the equation above as follows:
q p qv + ngasRT
The constant pressure heat of reaction is equal to the constant volume heat of reaction only when
there is no net production or no net consumption of gas. The ngas RT term that appears in these
equations is related to the amount of pressure-volume work that is done when the reaction is carried out
constant pressure: w = − Pex V = − P V = − ngasRT .
Example 1-9: A bomb calorimeter is heated electrically using a 15.00 watt heater. ◄ Note: 1 watt = 1 J s−1
It takes 568 seconds to raise the temperature of the calorimeter by 2.00 °C.
(a) What is the heat capacity of the calorimeter?
Formation reactions
In principle, we could use calorimetry to obtain r H for all reactions. ◄ There are many more chemical
However, it is neither practical nor desirable to measure r H for every reactions than there are substances.
Rather than measure and tabulate r H
possible reaction. It is more practical to obtain r H values for important
values for all possible reactions, it
reaction types. makes more sense to focus on reactions
A formation reaction is a reaction in which one mole of a substance is that produce a single compound.
formed from its elements in their reference states. With relatively few
exceptions, the reference state of an element is the most stable form of that
element at the temperature and pressure of interest.
Learning Tip: Make sure you
Here are the formation reactions for H2O(l), NH3(g), and Fe2O3(s).
can write down the formation
reaction for a given compound.
1
⎯⎯⎯
→ f H = −285.83 kJ mol−1
o
H2(g) + 2 O2(g) H2O(l)
1 3
H2(g) ⎯⎯⎯
→ f H = −45.94 kJ mol−1
o
2 N2(g) + 2 NH3(g)
3
O2(g) ⎯⎯⎯
→ f H = −824.2 kJ mol−1
o
2 Fe(s) + 2 Fe2O3(s)
f H is called the standard enthalpy of formation and in most tabulations, f H values are given for
T = 298.15 K. (See the table on the next page.) We will soon see that f H values can be used to
calculate r H for a specified reaction. Consequently, formation reactions are very important and
f H values are extremely useful.
Example 1-10: Draw Lewis structures for each of these molecular compounds and
then write formation reactions for each of them.
(a) FCHO(g) (b) H2SO4(l) (c) C6H5COOH(s)
CHEM 123/125 Module 1: Energy, Entropy, and Spontaneous Change page 33 of 42
Mass f H S Cp
o
Substance −1
/ g mol / kJ mol−1 / J K −1 mol−1 / J K −1 mol−1
Remarks:
• Δ f H is zero for an element in its reference form. This is because the formation reaction for an element involves
The initial and final states are identical, so we must have ∆H = Hf − Hi = 0 for the system and therefore, f H = 0 kJ
• Certain elements occur in more than one form. Carbon is an example you should remember.
Carbon occurs naturally as graphite or diamond. Graphite is the more stable form at 298.15 K and 1 bar. The
reference form of carbon is graphite.
o
• C p is the molar heat capacity at a constant pressure of 1 bar.
CHEM 123/125 Module 1: Energy, Entropy, and Spontaneous Change page 34 of 42
Consider the following general equation for a reaction involving the substances A,
B, C, and D.
aA + bB ⎯⎯⎯
→ cC + dD
Remember that the stoichiometric numbers have positive values for the products and
negative values for the reactants.
𝛥𝑟 𝐻° = ∑ 𝜈𝑘 × 𝛥𝑓 𝐻𝑘𝑜
⏟𝑘
𝑎 𝑤𝑒𝑖𝑔ℎ𝑡𝑒𝑑 𝑠𝑢𝑚
𝑜𝑓 𝛥𝑓 𝐻 𝑜 𝑣𝑎𝑙𝑢𝑒𝑠
𝑜 𝑜 𝑜 𝑜
= [𝑐
⏟ × 𝛥𝑓 𝐻𝐶 + 𝑑 × 𝛥𝑓 𝐻𝐷 + . . . ] − [𝑎
⏟ × 𝛥𝑓 𝐻𝐴 + 𝑏𝛥𝐻𝐵 + . . . ]
𝑎 𝑤𝑒𝑖𝑔ℎ𝑡𝑒𝑑 𝑠𝑢𝑚 𝑜𝑓 𝛥𝑓 𝐻° 𝑣𝑎𝑙𝑢𝑒𝑠 𝑎 𝑤𝑒𝑖𝑔ℎ𝑡𝑒𝑑 𝑠𝑢𝑚 𝑜𝑓 𝛥𝑓 𝐻° 𝑣𝑎𝑙𝑢𝑒𝑠
𝑓𝑜𝑟 𝑡ℎ𝑒 𝑝𝑟𝑜𝑑𝑢𝑐𝑡𝑠 𝑓𝑜𝑟 𝑡ℎ𝑒 𝑟𝑒𝑎𝑐𝑡𝑎𝑛𝑡𝑠
r H is the enthalpy change per mole of reaction for the following process.
Example 1-11: Use standard enthalpies of formation at 298.15 K to See problems 7-79 to 7-81,
7-84, 7-86 and 7-91 in
calculate r H for the following reaction. Petrucci (11th edition)
C6H6(l) + 15
2
O2(g) ⎯⎯⎯
→ 6 CO2(g) + 3 H2O(l)
CHEM 123/125 Module 1: Energy, Entropy, and Spontaneous Change page 36 of 42
Hess’ Law
The enthalpy change is the same regardless of whether the process is carried out in a single step or
via a sequence of steps.
As a result, we can use values of r H values for a given set of reactions to calculate r H for a new or
different reaction. In the following example, enthalpies of combustion are used together with Hess’ Law to
calculate the enthalpy of formation of a compound. This is one way to obtain enthalpies of formation of
compounds. (Enthalpies of combustion are relatively easy to obtain experimentally.)
Example 1-12: Use the thermochemical equations below to calculate f H for C2H6(g).
o
⎯⎯⎯
→ 2 CO2(g) + 3 H2O(l) r H = −1560 kJ mol−1
o
C2H6(g) + 7
2 O2(g)
CHEM 123/125 Module 1: Energy, Entropy, and Spontaneous Change page 37 of 42
The superscript o has a very precise interpretation in thermodynamics. Consider again the generalized
equation for a chemical reaction.
aA + bB ⎯⎯
→ cC + dD
The standard enthalpy of reaction is denoted by r HT . Strictly speaking, r HT is the enthalpy change per
o o
mole of reaction for the process of transforming a moles of pure A and b moles of pure B, each in its standard
state at temperature T, into c moles of pure C and d moles of pure D, each in its standard state at
temperature T. Thermodynamic standard states are defined as follows.
Gas Pure ideal gas at 1 bar ◄ For an ideal gas or ideal solution,
intermolecular forces are ignored.
Aqueous solution Ideal solution with concentration
of 1 mol L−1 at 1 bar.
It should be noted that the standard states of gases and aqueous solutions are idealized, “fictitious” states;
intermolecular forces are ignored.
As a result of the conventions adopted for the thermodynamic standard states, r H , is the standard
enthalpy of reaction for an idealized reaction that could never be carried out. We can explore this idea
further by focusing on the formation reaction for NH3(g).
This process is hypothetical because: (1) no gas behaves ideally at a pressure of 1 bar; and (2) we cannot
convert reactants into products while simultaneously maintaining the purity of each reactant. In order for the
reaction to occur, the reactants must be mixed together at some point.
It may seem strange that we should concern ourselves with the calculation of enthalpy changes (and eventually
entropy changes) for hypothetical processes involving pure unmixed substances. Herein lies the beauty and
power of thermodynamics: We can use standard thermodynamic properties of pure substances to calculate
thermodynamic properties of complex mixtures at any temperature and pressure we desire.
CHEM 123/125 Module 1: Energy, Entropy, and Spontaneous Change page 38 of 42
aA + bB ⎯⎯
→ cC + dD
We have already seen how to calculate r H for a reaction using f H values for the
𝛥𝑟 𝐻 𝑜 = ⏟
[ 𝑐𝛥𝑓 𝐻𝐶𝑜 + 𝑑𝛥𝑓 𝐻𝐷𝑜 + … ] − [ 𝑎𝛥𝑓 𝐻𝐴𝑜 + 𝑏𝛥𝑓 𝐻𝐵𝑜 + … ] = ∑ 𝜈𝑘 × 𝛥𝑓 𝐻𝑘𝑜
⏟
𝑎 𝑤𝑒𝑖𝑔ℎ𝑡𝑒𝑑 𝑠𝑢𝑚 𝑜𝑓 𝛥𝑓 𝐻° 𝑣𝑎𝑙𝑢𝑒𝑠 𝑎 𝑤𝑒𝑖𝑔ℎ𝑡𝑒𝑑 𝑠𝑢𝑚 𝑜𝑓 𝛥𝑓 𝐻° 𝑣𝑎𝑙𝑢𝑒𝑠 𝑘
𝑓𝑜𝑟 𝑡ℎ𝑒 𝑝𝑟𝑜𝑑𝑢𝑐𝑡𝑠 𝑓𝑜𝑟 𝑡ℎ𝑒 𝑟𝑒𝑎𝑐𝑡𝑎𝑛𝑡𝑠
In principle, we could calculate r S for a reaction using f S values for all ◄ f S represents the standard
entropy of formation. It is the entropy
of the substances involved. However, f S values are never tabulated. change per mole of reaction for a
reaction in which one mole of a
This is because the third law permits the determination of absolute
substance is formed from its elements
standard molar entropies, S , for substances. So, we use S values in their reference states at 1 bar.
We can calculate rG in two different ways. If r H and r S values are available,
r G = r H − T r S
In the next example, we use f H and S values for the reactants and products to calculate r H . r S ,
Example 1-13: For each of the following reactions, calculate r H , r S and rG at 298.15 K .
The Gibbs energy change for a process at constant temperature and pressure is given
by the following equation where all terms refer to changes for the system.
G = H − T S
rG is the Gibbs energy change per mole of reaction and is called the Gibbs energy of reaction.
• If rG is less than zero, the reaction proceeds spontaneously in the forward direction.
• If rG is greater than zero, the reaction proceeds spontaneously in the reverse direction.
rG < 0 the forward reaction is spontaneous; the reverse reaction is not allowed
rG > 0 the reverse reaction is spontaneous; the forward reaction is not allowed
Example 1-14: There are four cases that can be qualitatively analyzed based on the signs of the enthalpy
and entropy change. Use the criteria for spontaneous change and the equation for Gibbs energy change to
predict the sign of ∆G for each of the possible cases.
4
CHEM 123/125 Module 1: Energy, Entropy, and Spontaneous Change page 42 of 42
Keep in mind: Spontaneous does not mean fast. Consider the formation of graphite from diamond.
At 1 bar and 25 oC, the Gibbs energy of formation indicates the process is spontaneous, however, it is
extremely slow, and does not proceed on a timescale that humans would observe.
The summary below highlights some differences between thermodynamics and kinetics; two important
factors when discussing the progression of a chemical reaction.
• Extent of Reaction, how “far” the reaction will • Rate of reaction (rate constant, k)
go (thermodynamic equilibrium constant, K)
What determines rate? The rate of a chemical reaction (i.e. how fast it occurs) depends on the
magnitude of the energy barrier between the initial and transition states. You will learn more about this
concept in Module 2.
The reaction profile diagram below can help you visualize the idea of this energy barrier. The negative
value for 𝜟𝑮 indicates this process is spontaneous in the forward direction. The magnitude of the energy
barrier (𝛥𝐺 ‡) affects the rate of the reaction.
Reactants → Products
G ‡
reactants
G
products
reaction progress