Physical Chemistry 2
Physical Chemistry 2
Physical Chemistry 2
Patrick Fleming
California State University East Bay
California State University East Bay
Physical Chemistry
Patrick Fleming
This text is disseminated via the Open Education Resource (OER) LibreTexts Project (https://LibreTexts.org) and like the hundreds
of other texts available within this powerful platform, it is freely available for reading, printing and "consuming." Most, but not all,
pages in the library have licenses that may allow individuals to make changes, save, and print this book. Carefully
consult the applicable license(s) before pursuing such effects.
Instructors can adopt existing LibreTexts texts or Remix them to quickly build course-specific resources to meet the needs of their
students. Unlike traditional textbooks, LibreTexts’ web based origins allow powerful integration of advanced features and new
technologies to support learning.
The LibreTexts mission is to unite students, faculty and scholars in a cooperative effort to develop an easy-to-use online platform
for the construction, customization, and dissemination of OER content to reduce the burdens of unreasonable textbook costs to our
students and society. The LibreTexts project is a multi-institutional collaborative venture to develop the next generation of open-
access texts to improve postsecondary education at all levels of higher learning by developing an Open Access Resource
environment. The project currently consists of 14 independently operating and interconnected libraries that are constantly being
optimized by students, faculty, and outside experts to supplant conventional paper-based books. These free textbook alternatives are
organized within a central environment that is both vertically (from advance to basic level) and horizontally (across different fields)
integrated.
The LibreTexts libraries are Powered by NICE CXOne and are supported by the Department of Education Open Textbook Pilot
Project, the UC Davis Office of the Provost, the UC Davis Library, the California State University Affordable Learning Solutions
Program, and Merlot. This material is based upon work supported by the National Science Foundation under Grant No. 1246120,
1525057, and 1413739.
Any opinions, findings, and conclusions or recommendations expressed in this material are those of the author(s) and do not
necessarily reflect the views of the National Science Foundation nor the US Department of Education.
Have questions or comments? For information about adoptions or adaptions contact [email protected]. More information on our
activities can be found via Facebook (https://facebook.com/Libretexts), Twitter (https://twitter.com/libretexts), or our blog
(http://Blog.Libretexts.org).
This text was compiled on 07/10/2023
TABLE OF CONTENTS
Licensing
1: The Basics
1.1: The System and the Surroundings
1.2: Pressure and Molar Volume
1.3: Temperature
1.4: The Zeroth Law of Thermodynamics
1.5: Work and Energy
1.E: The Basics (Exercises)
1.S: The Basics (Summary)
2: Gases
2.1: The Empirical Gas Laws
2.2: The Ideal Gas Law
2.3: The Kinetic Molecular Theory of Gases
2.4: Kinetic Energy
2.5: Graham’s Law of Effusion
2.6: Collisions with Other Molecules
2.7: Real Gases
2.E: Gases (Exercises)
2.S: Gases (Summary)
1 https://chem.libretexts.org/@go/page/182391
5: The Second Law
5.1: Introduction to the Second Law
5.2: Heat Engines and the Carnot Cycle
5.3: Entropy
5.4: Calculating Entropy Changes
5.5: Comparing the System and the Surroundings
5.6: Entropy and Disorder
5.7: The Third Law of Thermodynamics
5.8: Adiabatic Compressibility
5.E: The Second Law (Exercises)
5.S: The Second Law (Summary)
8: Phase Equilibrium
8.1: Prelude to Phase Equilibrium
8.2: Single Component Phase Diagrams
8.3: Criterion for Phase Equilibrium
8.4: The Clapeyron Equation
8.5: The Clausius-Clapeyron Equation
8.6: Phase Diagrams for Binary Mixtures
8.7: Liquid-Vapor Systems - Raoult’s Law
8.8: Non-ideality - Henry's Law and Azeotropes
8.9: Solid-Liquid Systems - Eutectic Points
8.10: Cooling Curves
8.E: Phase Equilibrium (Exercises)
8.S: Phase Equilibrium (Summary)
2 https://chem.libretexts.org/@go/page/182391
9: Chemical Equilibria
9.1: Prelude to Chemical Equilibria
9.2: Chemical Potential
9.3: Activities and Fugacities
9.4: Pressure Dependence of Kp - Le Châtelier's Principle
9.5: Degree of Dissociation
9.6: Temperature Dependence of Equilibrium Constants - the van ’t Hoff Equation
9.7: The Dumas Bulb Method for Measuring Decomposition Equilibrium
9.8: Acid-Base Equilibria
9.9: Buffers
9.10: Solubility of Ionic Compounds
9.E: Chemical Equilibria (Exercises)
9.S: Chemical Equilibria (Summary)
10: Electrochemistry
10.1: Electricity
10.2: The connection to ΔG
10.3: Half Cells and Standard Reduction Potentials
10.4: Entropy of Electrochemical Cells
10.5: Concentration Cells
10.E: Electrochemistry (Exercises)
10.S: Electrochemistry (Summary)
3 https://chem.libretexts.org/@go/page/182391
12.E: Chemical Kinetics II (Exercises)
12.S: Chemical Kinetics II (Summary)
Index
Index
Glossary
Detailed Licensing
4 https://chem.libretexts.org/@go/page/182391
Licensing
A detailed breakdown of this resource's licensing can be found in Back Matter/Detailed Licensing.
1 https://chem.libretexts.org/@go/page/417261
CHAPTER OVERVIEW
1: The Basics
Thumbnail: Frying an egg is an example of a chemical change induced by the addition of thermal enegy (via heat). Image used
iwth permission (CC BY-SA 3.0; Managementboy).
This page titled 1: The Basics is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick Fleming.
1
1.1: The System and the Surroundings
The Zeroth Law of Thermodynamics deals with the temperature of a system. And while it may seem intuitive as to what terms like
“temperature” and “system” mean, it is important to define these terms. The easiest terms to define are the ones used to describe the
system of interest and the surroundings, both of which are subsets of the universe.
Universe – everything
System – subset of the universe that is being studied and/or measured
Surroundings – every part of the universe that is not the system itself.
As it turns out, there can be several types of systems, depending on the nature of the boundary that separates the system from the
surroundings, and specifically whether or not it allows to the transmittance of matter or energy across it.
Open System – allows for both mass and energy transfer across its boundary
Closed System – allows for energy transfer across its boundary, but not mass transfer
Isolated System – allows neither mass nor energy transfer across its boundary
Further, systems can be homogeneous (consisting of only a single phase of matter, and with uniform concentration of all
substances present throughout) or heterogeneous (containing multiple phases and/or varying concentrations of the constituents
throughout.) A very important variable that describes a system is its composition, which can be specified by the number of moles
of each component or the concentration of each component. The number of moles of a substance is given by the ratio of the number
of particles to Avogadro’s number
N
n =
NA
where n is the number of moles, N is the number of particles (atoms, molecules, or formula units) and N is Avogadro’s number
A
the ratio of volume and number of moles of substance. For a given substance, the molar volume is inversely proportional to the
density of the substance.
In a homogeneous system, an intensive variable will describe not just the system as a whole, but also any subset of that system.
However, this may not be the case in a heterogeneous system!
This page titled 1.1: The System and the Surroundings is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
1.1.1 https://chem.libretexts.org/@go/page/84288
1.2: Pressure and Molar Volume
Italian physicist Evangelista Torricelli (1608 – 1647) (Evangelista Torricelli) was the inventor of an ingenious device that could be
used to measure air pressure. Basically, he took a glass tube closed at one end, and filled it with mercury. He then inverted it,
submerging the open end below the surface level in a pool of mercury. The mercury in the glass tube was then allowed to drain,
leaving a vacuum (known as a “Torrocellian vacuum”) in the open space at the closed end of the tube.
Figure 1.2.1 : Evangelista Torricelli (1608 - 1647). Evangelista Torricelli portrayed on the frontpage of Lezioni d'Evangelista
Torricelli. (Public Domain).
Remarkably, the tube did not drain completely! Torricelli, was able to use the residual column height to measure the pressure of the
air pushing down on the surface of the pool of mercury. The larger the pressure pushing down on the exposed surface, the larger the
column height is observed to be. The ambient air pressure can be computed by equating the force generated by the mass of the
mercury in the column to the force generated by ambient air pressure (after normalizing for surface area). The resulting relationship
is
p = ρgh
where ρ is the density of the mercury (13.1 g/cm3), g is the acceleration due to gravity, and h is the height of the column. Torricelli
found that at sea level, the height of the column was 76 cm.
2 2 2
100 c m 1 kg 1kg m/s
3
p = (13.1 g/c m )(9.8 m/s)(76 cm) ( )( )( )
2
m 1000 g N
2 5
100, 000N / m = 1 × 10 P a
A force of 1 N acting on an area of 1 m2 is a Pascal (Pa). A standard atmosphere is 101,325 Pa (101.325 kPa), or 76.0 cm Hg (760
mm Hg.) Another commonly used unit of pressure is the bar:
1.2.1 https://chem.libretexts.org/@go/page/84289
Figure 1.2.2 : : Set up for the Torricelli barometer.
This page titled 1.2: Pressure and Molar Volume is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
1.2.2 https://chem.libretexts.org/@go/page/84289
1.3: Temperature
Another important variable that describes the state of a system it the system’s temperature. Like pressure, temperature scales
experienced an important process of development over time. Three of the most important temperature scales in US culture are the
Fahrenheit, Celsius, and Kelvin scales.
Figure 1.3.1: (left) G. Daniel Fahrenheit (1686 – 1736). (CC BY 4.0 ) and (right) Anders Celsius (1700 - 1780)
G. Daniel Fahrenheit wanted to develop a temperature scale that would be convenient to use in his laboratory. He wanted it to be of
convenient magnitude and wanted to avoid having to use any negative values for temperature. So he define the zero of his
temperature scale to be the lowest temperature he could create in his laboratory, which was in a saturated brine/water/ice slurry. He
then defined 100 °F as his own body temperature. As a result, using his temperature scale, water has a normal melting point (the
temperature at 1.00 atm pressure at which water ice melts) of 32 °F. Similarly, water boils (again at 1 atm pressure) at a temperature
of 212 °F. The difference between these values is 180 °F.
Anders Celsius also thought a 100 degree temperature scale made sense, and was given the name “the centigrade scale”. He defined
0 °C on his scale as the normal boiling point of water, and 100 °C as the normal freezing point. By today’s standards, this inverted
temperature scale makes little sense. The modern Celsius temperature scale defines 0 °C as the normal freezing point of water and
100 °C as the normal boiling point. The difference is 100 °C. Comparing this to the Fahrenheit scale, one can easily construct a
simple equation to convert between the two scales.
32°F = 0 °C (m) + b
b = 32 °F
9 °F
y °F = x °C ( ) + 32 °F
5 °C
1.3.1 https://chem.libretexts.org/@go/page/84290
5 °C
x °C = (y °F − 32 °F ) ( )
9 °F
Many physical properties of matter suggest that there is an absolute minimum temperature that can be attained by any sample. This
minimum temperature can be shown by several types or experiments to be -273.15 °C. An absolute temperature scale is one that
assigns the minimum temperature a value of 0. One particularly useful scale is named after William Lord Kelvin (Kelvin, Lord
William Thomson (1824-1907) , 2007).
1K
z K = x °C ( ) + 273.15 K
1 °C
This page titled 1.3: Temperature is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick Fleming.
1.3.2 https://chem.libretexts.org/@go/page/84290
1.4: The Zeroth Law of Thermodynamics
Temperature is an important property when it comes to measuring energy flow through a system. But how does one use or
measure temperature? Fortunately, there is a simple and intuitive relationship which can be used to design a thermometer – a device
to be used to measure temperature and temperature changes. The zeroth law of thermodynamics can be stated as follows:
This page titled 1.4: The Zeroth Law of Thermodynamics is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated
by Patrick Fleming.
1.4.1 https://chem.libretexts.org/@go/page/84291
1.5: Work and Energy
Temperature, pressure and volume are important variables in the description of physical systems. They will also be important to
describe how energy flows from one system to another. Generally, energy can flow in two important forms: 1) work and 2) heat.
The bookkeeping needed to track the flow of energy is what the subject of Thermodynamics is all about, so these topics will be
discussed at length in subsequent chapters. However, a little bit of review is in order, just to set the foundation for the discussions
that are forthcoming.
Energy
Energy is an important entity in the modern world. We use energy to light our homes, drive our cars, and power our electronic
devices. According to Richard Smalley, co-winner of the 1996 Nobel Prize in Chemistry, energy is one of the (if not the biggest)
challenge we face moving into the 21st century (energy.senate.gov, 2004):
Energy is at the core of virtually every problem facing humanity. We cannot afford to get
this wrong. … Somehow we must find the basis for energy prosperity for ourselves and the
rest of humanity for the 21st century. By the middle of this century we should assume we
will need to at least double world energy production from its current level, with most of
this coming from some clean, sustainable, CO2-free source. For worldwide peace and
prosperity it needs to be cheap.- Richard Smalley, Testimony to the Senate Committee on
Energy and Natural Resources, April 26, 2004
Energy can be measured in a multitude of different units, including joules (J), kilojoules (kJ), calories (cal), kilocalories (kcal), as
well as several other set of units such as kJ/mol or kcal/mol.
A calorie (cal) was once defined as the amount of energy needed to raise the temperature of 1 g of water by 1 °C. This definition
suggests a convenient property of water called the specific heat:
1 cal
C =
g °C
where a joule is the energy necessary to move a mass a distance of 1 m against a resisting force of 1 N.
A dietary Calorie (Cal) is equal to 1000 cal, or 1 kcal, and is often listed on the labels of
food containers to indicate the energy content of the food inside.
Energy can take the form of potential energy (stored energy) and kinetic energy (realized energy) forms. Kinetic energy is the
energy of motion. On the other hand, potential energy can be defined as the energy stored in a system that can be converted to
kinetic energy someplace in the universe. Kinetic energy of a particle can be expressed as
1 2
Ekin = mv (1.5.1)
2
where m is the mass of the particle, and v is the magnitude of its velocity (or speed). Equation 1.5.1 describes the kinetic energy
associated with translation; other expressions exist for different motions (e.g., rotation or vibration).
An example of a system in which energy is converted between kinetic energy and potential energy is a Hooke’s Law oscillator.
According to Hooke’s Law, the force acting on an object is proportional in magnitude to the displacement of the object from an
equilibrium position, and opposite in sign.
F = −kx (1.5.2)
In this equation, F is the force, x is the displacement from equilibrium, and k is the constant of proportionality. The negative sign
is necessary to insure that the force acting on the object is one that will tend to restore it to an equilibrium position (x = 0 )
1.5.1 https://chem.libretexts.org/@go/page/84292
irrespective of whether x is positive or negative.
As the object that follows Hooke’s Law moves from its equilibrium position, the kinetic energy of its motion is converted into
potential energy until there is no more kinetic energy left. At this point, the change in displacement will change direction, returning
the object to the equilibrium position by converting potential energy back into kinetic energy.
U (x) = − ∫ (−kx)dx
1
2
= kx + constant
2
With the proper choice of coordinate system and other definitions, the constant of integration can be arbitrarily made to be zero (for
example, by choosing it to offset any other forces acting on the object, such as the force due to gravity.) The kinetic energy is then
given by the total energy minus the potential energy (since the total energy must be constant due to the conservation of energy in
the system!)
Work
Work is defined as the amount of energy expended to move a mass against a resisting force. For a mass being moved along a
surface, the amount of energy expended must be sufficient to overcome the resisting force (perhaps due to friction) and also
sufficient to cause motion along the entire path.
1.5.2 https://chem.libretexts.org/@go/page/84292
The energy expended as work in this case (if the force is independent of the position of the object being moved) is given by
w = −F Δx
where F is the magnitude of the resisting force, and Dx is the displacement of the object. The negative sign is necessary since the
force is acting in the opposite direction of the motion. A more general expression, and one that can be used if the force is not
constant over the entire motion is
dw = −F dx
This expression can then be integrated, including any dependence F might have on x as needed for a given system.
Another important way that work can be defined includes that for the expansion of a gas sample against an external pressure. In
this case, the displacement is defined by a change in volume for the sample:
dw = −pext dV
This is a very convenient expression and will be used quite often when discussing the work expended in the expansion of a gas.
The conversion of potential energy into kinetic energy generally is accomplished through work which is done someplace in the
universe. As such, the concepts of energy and work are inexorably intertwined. They will be central to the study of
thermodynamics.
pint = pext = p
In this case, the work of expansion can be calculated by integrating the expression for dw.
w =∫ dw = − ∫ pdV
allows for the expression in terms of volume and temperature. If the temperature is constant (so that it can be placed before the
integral) the expression becomes
V2
dV V2
w = −nRT ∫ = −nRT ln( ) (1.5.5)
V1
V V1
where V and V are the initial and final volumes of the expansion respectively.
1 2
Example 1.5.1:
Consider 1.00 mol of an ideal gas, expanding isothermally at 273 K, from an initial volume of 11.2 L to a final volume of 22.4
L. What is the final pressure of the gas? Calculate the work of the expansion if it occurs
a. against a constant external pressure equal to the final pressure you have calculated.
b. reversibly.
Solution
First, let’s calculate the final pressure via Equation 1.5.4:
−1 −1
nRT (1.00 mole)(0.08206 atm L mol K )(273 K)
p = = = 1.00 atm
V 22.4 L
1.5.3 https://chem.libretexts.org/@go/page/84292
(This may be a relationship you remember from General Chemistry – that 1 mole of an idea gas occupies 22.4 L at 0 oC!)
Okay – now for the irreversible expansion against a constant external pressure:
dw = −pext dV
so
V2
w = −pexp ∫ dV = −pext ΔV
V1
But what the heck is an atm L? It is actually a fairly simply thing to convert from units of atm L to J by using the ideal gas law
constant.
J
8.314
mol K
w = −(11.2 atm L) ( ) = −1130 J
atm L
0.08206
mol K
Note that the negative sign indicated that the system is expending energy by doing work on the surroundings. (This concept
will be vital in the Chapter 3!)
Now for the reversible pathway. The work done by the system can be calculated for this change using Equation 1.5.5:
−1 −1
22.4 L
w = −(1.00 mol)(8.314J mol K )(273 K) ln( ) = −1570 J
11.2 L
Notes:
First notice how the value for the gas law constant, R, was chosen in order to match the units required in the problem.
(Read and recite that previous sentence to yourself a few times. The incorrect choice of the value of R is one of the most
common errors made by students in physical chemistry! By learning how the units will dictate your choice of R, you will
save yourself a considerable number of headaches as you learn physical chemistry!)
Second, You may note that the magnitude of work done by the system in the reversible expansion is larger than that of the
irreversible expansion. This will always be the case!
[1] There are many cases of “limiting ideal behavior” which we use to derive and/or explore the nature of chemical systems. The
most obvious case, perhaps, is that of the Ideal Gas Law.
This page titled 1.5: Work and Energy is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
1.5.4 https://chem.libretexts.org/@go/page/84292
1.E: The Basics (Exercises)
Q1.1
Convert the temperatures indicated to complete the following table
°F °C K
25
98.6
373.15
-40
32
Q1.2
Make a graph representing the potential energy of a harmonic oscillator as a function of displacement from equilibrium. On the
same graph, include a function describing the kinetic energy as a function of displacement from equilibrium as well as the total
energy of the system.
Q1.3
Calculate the work required to move a 3.2 kg mass 10.0 m against a resistive force of 9.80 N.
Q1.4
Calculate the work needed for a 22.4 L sample of gas to expand to 44.8 L against a constant external pressure of 0.500 atm.
Q1.5
If the internal and external pressure of an expanding gas are equal at all points along the entire expansion pathway, the expansion is
called “reversible.” Calculate the work of a reversible expansion for 1.00 mol of an ideal gas expanding from 22.4 L at 273 K to a
final volume of 44.8 L.
This page titled 1.E: The Basics (Exercises) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
1.E.1 https://chem.libretexts.org/@go/page/84293
1.S: The Basics (Summary)
Learning Objectives
Upon mastering the material covered in this chapter, one should be able to do the following:
1. Write down expressions from which work of motion and of expansion can be calculated.
2. Express the “Zeroth Law of Thermodynamics.
3. Convert between temperatures on several scales that are commonly used.
4. Define boundaries that differentiate between a system and its surroundings.
5. Perform calculations involving Specific Heat and understand how the specific heat governs temperature changes for the flow of
a given amount of energy.
References
1. Biography.com. (n.d.). Anders Celcius. (A&E Television Networks) Retrieved March 10, 2016, from Biography.com:
http://www.biography.com/people/ande...elsius-9242754
2. BIPM. (n.d.). International Committee for Weights and Measures (CIPM). Retrieved March 10, 2016, from BIPM: Bereau
International des Poids et Mesures: http://www.bipm.org/en/committees/cipm/
3. energy.senate.gov. (2004, April 26). Testimony of R. E. Smalley to the Senate Committee on Energy and Natural Resources;
Hearing on sustainable , low emission, elect. Retrieved March 10, 2016, from Energy Bulletin:
www2.energybulletin.net/node/249
4. Evangelista Torricelli. (n.d.). Retrieved March 3, 2016, from Famous Scientists: http://www.famousscientists.org/evan...ta-
torricelli/
1.S.1 https://chem.libretexts.org/@go/page/84436
5. Gabriel Fahrenheit Biography. (2016). Retrieved March 10, 2016, from Encyclopedia of World Biography:
http://www.notablebiographies.com/Du...t-Gabriel.html
6. Kelvin, Lord William Thomson (1824-1907) . (2007). (Wolfram Research) Retrieved March 10, 2016, from
scienceworld.wolfram.com: http://scienceworld.wolfram.com/biography/Kelvin.html
7. Mangum, B. W., & Furukawa, G. T. (1990). Guidelines for Realizing the International Practical Temperature Scale of 1990
(ITS-90), NIST Technical Note 1265. Gaithersberg, MD: National Institutes of Standards and Technology.
8. Strouse, G. F. (2008, January). Standard Platinum Resistance Thermometer Calibrations from the Ar TP to the Ag FP. National
Institute of Standards and Technology Special Publication 250-81.
This page titled 1.S: The Basics (Summary) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
1.S.2 https://chem.libretexts.org/@go/page/84436
CHAPTER OVERVIEW
2: Gases
This page titled 2: Gases is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick Fleming.
1
2.1: The Empirical Gas Laws
A number of important relationships describing the nature of gas samples have been derived completely empirically (meaning
based solely on observation rather making an attempt to define the theoretical reason these relationships may exist. These are the
empirical gas laws.
Boyle’s Law
One of the important relationships governing gas samples that can be modeled mathematically is the relationship between pressure
and volume. Robert Boyle (1627 – 1691) (Hunter, 2004) did experiments to confirm the observations of Richard Towneley and
Henry Powers to show that for a fixed sample of gas at a constant temperature, pressure and volume are inversely proportional.
pV = constant
or
p1 V2 = p2 V2
Boyle used a glass u-tube that was closed at one end and with the lower portion filled with mercury (trapping a sample of air in the
closed end.) By adding mercury to the open end, he was able to observe and quantify the compression of the trapped air.
Figure 2.1.1 : An apparatus similar to that used by Robert Boyle (1627 - 1691). Image taken from (Fazio, 1992).
Charles’ Law
Charles’ Law states that the volume of a fixed sample of gas at constant pressure is proportional to the temperature. For this law to
work, there must be an absolute minimum to the temperature scale since there is certainly an absolute minimum to the volume
scale!
V
= constant
T
or
V1 V1
=
T2 T2
The second law of thermodynamics also predicts an absolute minimum temperature, but that will be developed in a later chapter.
Gay-Lussac’s Law
Gay-Lussac’s Law states that the pressure of a fixed sample of gas is proportional to the temperature. As with Charles’ Law, this
suggests the existence of an absolute minimum to the temperature scale since the pressure can never be negative.
2.1.1 https://chem.libretexts.org/@go/page/84295
p
= constant
T
or
p1 p1
=
T2 T2
or
p1 V1 p2 V2
=
T1 T2
Avogadro’s Law
Amedeo Avogadro (1776-1856) (Encycolopedia, 2016) did extensive work with gases in his studies of matter. In the course of his
work, he noted an important relationship between the number of moles in a gas sample. Avogadro’s Law (Avogadro, 1811) states
that at the same temperature and pressure, any sample of gas has the same number of molecules per unit volume.
n
= constant
V
or
n1 n2
=
V1 V2
This page titled 2.1: The Empirical Gas Laws is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
2.1.2 https://chem.libretexts.org/@go/page/84295
2.2: The Ideal Gas Law
The ideal gas law combines the empirical laws into a single expression. It also predicts the existence of a single, universal gas
constant, which turns out to be one of the most important fundamental constants in science.
pV = nRT
The ideal gas law constant is of fundamental importance and can be expressed in a number of different sets of units.
Value Units
The ideal gas law, as derived here, is based entirely on empirical data. It represents “limiting ideal behavior.” As such, deviations
from the behavior suggested by the ideal gas law can be understood in terms of what conditions are required for ideal behavior to
be followed (or at least approached.) As such, it would be nice if there was a theory of gases that would suggest the form of the
ideal gas law and also the value of the gas law constant. As it turns out, the kinetic molecular theory of gases does just that!
This page titled 2.2: The Ideal Gas Law is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
2.2.1 https://chem.libretexts.org/@go/page/84296
2.3: The Kinetic Molecular Theory of Gases
Theoretical models attempting to describe the nature of gases date back to the earliest scientific inquiries into the nature of matter
and even earlier! In about 50 BC, Lucretius, a Roman philosopher, proposed that macroscopic bodies were composed of atoms that
continually collide with one another and are in constant motion, despite the observable reality that the body itself is as rest.
However, Lucretius’ ideas went largely ignored as they deviated from those of Aristotle, whose views were more widely accepted
at the time.
In 1738, Daniel Bernoulli (Bernoulli, 1738) published a model that contains the basic framework for the modern Kinetic Molecular
theory. Rudolf Clausius furthered the model in 1857 by (among other things) introducing the concept of mean free path (Clausius,
1857). These ideas were further developed by Maxwell (Maxwell, Molecules, 1873). But, because atomic theory was not fully
embraced in the early 20th century, it was not until Albert Einstein published one of his seminal works describing Brownian motion
(Einstein, 1905) in which he modeled matter using a kinetic theory of molecules that the idea of an atomic (or molecular) picture
really took hold in the scientific community.
In its modern form, the Kinetic Molecular Theory of gasses is based on five basic postulates.
1. Gas particles obey Newton’s laws of motion and travel in straight lines unless they collide with other particles or the walls of
the container.
2. Gas particles are very small compared to the averages of the distances between them.
3. Molecular collisions are perfectly elastic so that kinetic energy is conserved.
4. Gas particles so not interact with other particles except through collisions. There are no attractive or repulsive forces between
particles.
5. The average kinetic energy of the particles in a sample of gas is proportional to the temperature.
Qualitatively, this model predicts the form of the ideal gas law.
1. More particles means more collisions with the wall (\(p \propto n\))
2. Smaller volume means more frequent collisions with the wall (\(p \propto 1/V\))
3. Higher molecular speeds means more frequent collisions with the walls (\(p \propto T\))
Putting all of these together yields
nT nT
p ∝ =k
V V
which is exactly the form of the ideal gas law! The remainder of the job is to derive a value for the constant of proportionality (k )
that is consistent with experimental observation.
For simplicity, imagine a collection of gas particles in a fixed-volume container with all of the particles traveling at the same
velocity. What implications would the kinetic molecular theory have on such a sample? One approach to answering this question is
to derive an expression for the pressure of the gas.
The pressure is going to be determined by considering the collisions of gas molecules with the wall of the container. Each collision
will impart some force. So the greater the number of collisions, the greater the pressure will be. Also, the larger force imparted per
collision, the greater the pressure will be. And finally, the larger the area over which collisions are spread, the smaller the pressure
will be.
(number of collisions) × (force imparted per collision)
p ∝
area
2.3.1 https://chem.libretexts.org/@go/page/84297
Figure 2.3.1 : The "collision volume" is the subset of the total volume that contains molecules that will actually collide with area A
in the time interval Δt.
First off, the pressure that the gas exerts on the walls of the container would be due entirely to the force imparted each time a
molecule collides with the interior surface of the container. This force will be scaled by the number of molecules that hit the area of
the wall in a given time. For this reason, it is convenient to define a “collision volume”.
where v is the speed the molecules are traveling in the x direction, \(\Delta t\) is the time interval (the product of v ⋅ ΔT gives
x x
the length to the collision volume box) and A is the area of the wall with which the molecules will collide. Half of the molecules
within this volume will collide with the wall since half will be traveling toward it and half will be traveling away from it. The
number of molecules in this collision volume will be given by the total number of molecules in the sample and the fraction of the
total volume that is the collision volume. And thus, the number of molecules that will collide with the wall is given by
1 Vcol
Ncol = Ntot
2 V
And thus the number of molecules colliding with the wall will be
1 (vx ⋅ Δt) ⋅ A
Ncol = Ntot
2 V
The magnitude of that force imparted per collision will be determined by the time-rate of change in momentum of each particle as
it hits the surface. It can be calculated by determining the total momentum change and dividing by the total time required for the
event. Since each colliding molecule will change its velocity from vx to –vx, the magnitude of the momentum change is 2(mvx).
Thus the force imparted per collision is given by
2(m vx )
F =
Δt
2
mvx
= Ntot ( )A (2.3.3)
V
Since the pressure is given as the total force exerted per unit area, the pressure is given by
2
Ftot mvx Ntot m
2
p = = Ntot ( ) = vx
A V V
The question then becomes how to deal with the velocity term. Initially, it was assumed that all of the molecules had the same
velocity, and so the magnitude of the velocity in the x-direction was merely a function of the trajectory. However, real samples of
2.3.2 https://chem.libretexts.org/@go/page/84297
gases comprise molecules with an entire distribution of molecular speeds and trajectories. To deal with this distribution of values,
we replace (v ) with the squared average of velocity in the x direction ⟨v ⟩ .
2
x x
2
Ntot m 2
p = ⟨vx ⟩ (2.3.4)
V
The distribution function for velocities in the x direction, known as the Maxwell-Boltzmann distribution, is given by:
−−−−−− 2
m −mvx
f (vx ) = √ exp( )
2π kB T 2 kB T
normalization term exponential term
This function has two parts: a normalization constant and an exponential term. The normalization constant is derived by noting
that
∞
The Maxwell-Boltzmann distribution has to be normalized because it is a continuous probability distribution. As such, the sum
of the probabilities for all possible values of vx must be unity. And since v x can take any value between -∞ and ∞, then
Equation 2.3.5 must be true. So if the form of f (v ) is assumed to be
x
2
mvx
f (vx ) = N exp − ( )
2 kB T
The expression can be simplified by letting α = m/2k BT . It is then more simply written
∞ 2
−mvx
N ∫ exp( )dvx = 1
−∞
2 kB T
So
−− 1/2
π m
N√ =( )
α 2π kB T
Calculating an average for a finite set of data is fairly easy. The average is calculated by
N
1
x̄ = ∑ xi
N
i=1
But how does one proceed when the set of data is infinite? Or how does one proceed when all one knows are the probabilities
for each possible measured outcome? It turns out that that is fairly simple too!
2.3.3 https://chem.libretexts.org/@go/page/84297
N
x̄ = ∑ xi Pi
i=1
where P is the probability of measuring the value x . This can also be extended to problems where the measurable properties
i i
are not discrete (like the numbers that result from rolling a pair of dice) but rather come from a continuous parent population.
In this case, if the probability is of measuring a specific outcome, the average value can then be determined by
x̄ = ∫ xP (x)dx
where P (x) is the function describing the probability distribution, and with the integration taking place across all possible
values that x can take.
A value that is useful (and will be used in further developments) is the average velocity in the x direction. This can be derived
using the probability distribution, as shown in the mathematical development box above. The average value of v is given by x
This integral will, by necessity, be zero. This must be the case as the distribution is symmetric, so that half of the molecules are
traveling in the +x direction, and half in the –x direction. These motions will have to cancel. So, a more satisfying result will be
given by considering the magnitude of v , which gives the speed in the x direction. Since this cannot be negative, and given
x
⟨| vx |⟩ = 2 ∫ vx (f (vx )dx
0
In other words, we will consider only half of the distribution, and then double the result to account for the half we ignored.
For simplicity, we will write the distribution function as
2
f (vx ) = N exp(−α vx )
where
1/2
m
N =( )
2π kB T
and
m
α = .
2 kB T
so
1 N
⟨vx ⟩ = 2N ( ) =
2α α
This expression indicates the average speed for motion of in one direction.
2.3.4 https://chem.libretexts.org/@go/page/84297
However, real gas samples have molecules not only with a distribution of molecular speeds and but also a random distribution of
directions. Using normal vector magnitude properties (or simply using the Pythagorean Theorem), it can be seen that
2 2 2 2
⟨v⟩ = ⟨vx ⟩ + ⟨vy ⟩ + ⟨vz ⟩
Since the direction of travel is random, the velocity can have any component in x, y, or z directions with equal probability. As such,
the average value of the x, y, or z components of velocity should be the same. And so
2 2
⟨v⟩ = 3⟨vx ⟩
Substituting this into the expression for pressure (Equation 2.3.4) yields
Ntot m 2
p = ⟨v⟩
3V
All that remains is to determine the form of the distribution of velocity magnitudes the gas molecules can take. One of the first
people to address this distribution was James Clerk Maxwell (1831-1879). In his 1860 paper (Maxwell, Illustrations of the
dynamical theory of gases. Part 1. On the motions and collisions of perfectly elastic spheres, 1860), proposed a form for this
distribution of speeds which proved to be consistent with observed properties of gases (such as their viscosities). He derived this
expression based on a transformation of coordinate system from Cartesian coordinates (x, y , z ) to spherical polar coordinates (v , θ ,
ϕ ). In this new coordinate system, v represents the magnitude of the velocity (or the speed) and all of the directional data is carried
Applying this transformation of coordinates, and ignoring the angular part (since he was interested only in the speed) Maxwell’s
distribution (Equation 2.3.6) took the following form
2
2
mv
f (v) = N v exp ( ) (2.3.7)
2 kB T
This function has three basic parts to it: a normalization constant (N ), a velocity dependence (v ), and an exponential term that
2
contains the kinetic energy (½mv ). Since the function represents the fraction of molecules with the speed v , the sum of the
2
fractions for all possible velocities must be unity. This sum can be calculated as an integral. The normalization constant ensures that
∞
∫ f (v)dv = 1
0
At low velocities, the v term causes the function to increase with increasing v , but then at larger values of v , the exponential term
2
causes it to drop back down asymptotically to zero. The distribution will spread over a larger range of speed at higher temperatures,
but collapse to a smaller range of values at lower temperatures (Table 2.3.1).
2.3.5 https://chem.libretexts.org/@go/page/84297
Figure 2.3.1: Maxwell Distribution of speeds for hydrogen molecules at differing temperatures.
Using the Maxwell distribution as a distribution of probabilities, the average molecular speed in a sample of gas molecules can
be determined.
∞
−−−−−−−−−−
3 ∞ 2
m 3
mv
= 4π √( ) ∫ v exp ( ) dv (2.3.11)
2π kB T −∞
2 kB T
So
⎡ ⎤
−−−−−−−−−−
3
m ⎢ 1 ⎥
⟨v⟩ = 4π √( ) ⎢ ⎥
⎢ 2 ⎥
2π kB T ⎢ m ⎥
2( )
⎣ ⎦
2 kB T
Which simplifies to
1/2
8 kB T
⟨v⟩ = ( )
πm
Note: the value of ⟨v⟩ is twice that of ⟨v x⟩ which was derived in an earlier example!
⟨v⟩ = 2⟨vx ⟩
Example 2.3.1:
What is the average value of the squared speed according to the Maxwell distribution law?
Solution
2.3.6 https://chem.libretexts.org/@go/page/84297
∞
2 2
⟨v ⟩ = ∫ v f (v)dv (2.3.12)
−∞
−−−−−−−−−−
∞ 3 2
2
m 2
mv
=∫ v 4π √( ) v exp ( ) dv (2.3.13)
−∞
2π kB T 2 kB T
−−−−−−−−−−
3 ∞ 2
m 4
mv
= 4π √( ) ∫ v exp ( ) dv (2.3.14)
2π kB T −∞
2 kB T
⎡ ⎤
−−−−−−−−−−
3
⎢ −−−−−−− −⎥
m 1⋅3 π
2 ⎢ ⎥
⟨v ⟩ = 4π √( ) ⎢ ⎥
⎢ 2 ⎥
2π kB T m m
⎢ 3 ⎥
2 ( ) ( )
⎷
⎣ 2 kB T 2 kB T ⎦
which simplifies to
2
3 kB T
⟨v ⟩ =
m
Note: The square root of this average squared speed is called the root mean square (RMS) speed, and has the value
1/2
−−− 3 kB T
2
vrms = √⟨v ⟩ = ( )
m
The entire distribution is also affected by molecular mass. For lighter molecules, the distribution is spread across a broader range of
speeds at a given temperature, but collapses to a smaller range for heavier molecules (Table 2.3.2).
Figure 2.3.2: Maxwell Distribution of speeds at 800 K for different gasses of differing molecular masses.
The probability distribution function can also be used to derive an expression for the most probable speed (v ), the average (v ), mp ave
and the root-mean-square (v ) speeds as a function of the temperature and masses of the molecules in the sample. The most
rms
probable speed is the one with the maximum probability. That will be the speed that yields the maximum value of f (v) . It is
found by solving the expression
d
f (v) = 0
dv
2.3.7 https://chem.libretexts.org/@go/page/84297
for the value of v that makes it true. This will be the value that gives the maximum value of f (v) for the given temperature.
Similarly, the average value can be found using the distribution in the following fashion
vave = ⟨v⟩
and the root-mean-square (RMS) speed by finding the square root of the average value of v . Both demonstrated above.
2
−−−
2
vrms = √⟨v ⟩
This page titled 2.3: The Kinetic Molecular Theory of Gases is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by Patrick Fleming.
2.3.8 https://chem.libretexts.org/@go/page/84297
2.4: Kinetic Energy
Using expressions for v mp ,v
ave , or v
rms , it is fairly simple to derive expressions for kinetic energy from the expression
1
2
Ekin = mv
2
It is important to remember that there will be a full distribution of molecular speeds in a thermalized sample of gas. Some
molecules will be traveling faster and some more slowly. It is also important to recognize that the most probable, average, and
RMS kinetic energy terms that can be derived from the Kinetic Molecular Theory do not depend on the mass of the molecules
(Table 2.4.1). As such, it can be concluded that the average kinetic energy of the molecules in a thermalized sample of gas depends
only on the temperature. However, the average speed depends on the molecular mass. So, for a given temperature, light molecules
will travel faster on average than heavier molecules.
Table 2.4.1: Kinetic Properties of a Thermalized Ensemble (i.e., follows Maxwell-Boltzmann Distribution)
Property Speed Kinetic Energy
−−−− −
2 kb T
Most probable √ kB T
m
−−−− −
8 kb T 4 kB T
Average √
πm π
−−−− −
3 kb T 3
Root-mean-square √ kB T
m 2
Replacing ⟨v⟩ with the square of the RMS speed expression yields
2
Ntot m 3 kB T
p = ( )
3V m
which simplifies to
Ntot kB T
p =
V
Noting that Ntot = n·NA, where n is the number of moles and NA is Avogadro’s number
nNA kB T
p =
V
or
pV = nNA kB T
pV = nRT
That’s kind of cool, no? The only assumptions (beyond the postulates of the Kinetic Molecular Theory) is that the distribution of
velocities for a thermalized sample of gas is described by the Maxwell-Boltzmann distribution law. The next development will be
to use the Kinetic Molecular Theory to describe molecular collisions (which are essential events in many chemical reactions.)
2.4.1 https://chem.libretexts.org/@go/page/84298
Collisions with the Wall
In the derivation of an expression for the pressure of a gas, it is useful to consider the frequency with which gas molecules collide
with the walls of the container. To derive this expression, consider the expression for the “collision volume”.
Vcol = vx Δt ⋅ A
All of the molecules within this volume, and with a velocity such that the x-component exceeds vx (and is positive) will collide
with the wall. That fraction of molecules is given by
N ⟨v⟩Δt ⋅ A
Ncol =
V 2
and the frequency of collisions with the wall per unit area per unit time is given by
N ⟨v⟩
Zw =
V 2
In order to expand this model into a more useful form, one must consider motion in all three dimensions. Considering that
−−−−−−−−−−−−−−
⟨v⟩ = √ ⟨vx ⟩ + ⟨vy ⟩ + ⟨vz ⟩
and that
⟨v⟩ = 2⟨vx ⟩
or
1
⟨vx ⟩ = ⟨v⟩
2
and so
1 N
Zw = ⟨v⟩
4 V
The factor of N/V is often referred to as the “number density” as it gives the number of molecules per unit volume. At 1 atm
pressure and 298 K, the number density for an ideal gas is approximately 2.5 x 1019 molecule/cm3. (This value is easily calculated
using the ideal gas law.) By comparison, the average number density for the universe is approximately 1 molecule/cm3.
This page titled 2.4: Kinetic Energy is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick Fleming.
2.4.2 https://chem.libretexts.org/@go/page/84298
2.5: Graham’s Law of Effusion
An important consequence of the kinetic molecular theory is what it predicts in terms of effusion and diffusion effects. Effusion is
defined as a loss of material across a boundary. A common example of effusion is the loss of gas inside of a balloon over time.
The rate at which gases will effuse from a balloon is affected by a number of factors. But one of the most important is the
frequency with which molecules collide with the interior surface of the balloon. Since this is a function of the average molecular
speed, it has an inverse dependence on the square root of the molecular weight.
1
Rate of effusion ∝ −−−−
√M W
This can be used to compare the relative rates of effusion for gases of different molar masses.
\
Figure 2.5.1 : Effusion of gas particles through an orifice. (CC BY-SA 3.0; Astrang13).
This makes a convenient arrangement to measure the vapor pressure of the material inside the cell, as the total mass lost by
effusion through the orifice will be proportional to the vapor pressure of the substance. The vapor pressure can be related to the
mass lost by the expression
−−−−−−
g 2πRT
p = √
AΔt MW
where g is the mass lost, A is the area of the orifice, Δt is the time the effusion is allowed to proceed, T is the temperature and
M W is the molar mass of the compound in the vapor phase. The pressure is then given by p . A schematic of what a Knudsen
2.5.1 https://chem.libretexts.org/@go/page/84299
This page titled 2.5: Graham’s Law of Effusion is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
2.5.2 https://chem.libretexts.org/@go/page/84299
2.6: Collisions with Other Molecules
A major concern in the design of many experiments is collisions of gas molecules with other molecules in the gas phase. For
example, molecular beam experiments are often dependent on a lack of molecular collisions in the beam that could degrade the
nature of the molecules in the beam through chemical reactions or simply being knocked out of the beam.
In order to predict the frequency of molecular collisions, it is useful to first define the conditions under which collisions will occur.
For convenience, consider all of the molecules to be spherical and in fixed in position except for one which is allowed to move
through a “sea” of other molecules. A molecular collision will occur every time the center of the moving molecule comes within
one molecular diameter of the center of another molecule.
One can easily determine the number of molecules the moving molecule will “hit” by determining the number of molecules that lie
within the “collision cylinder”. Because we fixed the positions of all but one of the molecules, we must use the relative speed of the
moving molecule, which will be given by
–
vrel = √2 × v
The collisional cross section, which determined by the size of the molecule is given by
2
σ = πd
He 0.21
Ne 0.24
N2 0.43
CO2 0.52
C2H4 0.64
2.6.1 https://chem.libretexts.org/@go/page/84300
the number of collisions is given by
p –
Ncol = (√2 vΔtσ)
kB T
The frequency of collisions (number of collisions per unit time) is then given by
–
√2pσ
Z = ⟨v⟩
kB T
Perhaps a more useful value is the mean free path (λ ), which is the distance a molecule can travel on average before it collides
with another molecule. This is easily derived from the collision frequency. How far something can travel between collisions is
given by the ratio of how fast it is traveling and how often it hits other molecules:
⟨v⟩
λ =
Z
The mere fact that molecules undergo collisions represents a deviation from the kinetic molecular theory. For example, if molecules
were infinitesimally small (σ ≈ 0 ) then the mean free path would be infinitely long! The finite size of molecules represents one
significant deviation from ideality. Another important deviation stems from the fact that molecules do exhibit attractive and
repulsive forces between one another. These forces depend on a number of parameters, such as the distance between molecules and
the temperature (or average kinetic energy of the molecules.)
This page titled 2.6: Collisions with Other Molecules is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
2.6.2 https://chem.libretexts.org/@go/page/84300
2.7: Real Gases
While the ideal gas law is sufficient for the prediction of large numbers of properties and behaviors for gases, there are a number of
times that deviations from ideality are extremely important.
or
RT a
p = − (2.7.2)
2
Vm − b Vm
In this expression, a and b are variables of a given substance which can be measured and tabulated. In general, molecules with
large intermolecular forces will have large values of a , and large molecules will have large values of b. Some van der Waals
constants are given in Table 2.7.1.
The Table 2.7.1 : van der Waals constants for select Species
van Gas frequency of collisions a (atm L2 mol-2) b (L/mol)
der
He 0.0341 0.0238
Wal
ls N2 1.352 0.0387
mo CO2 3.610 0.0429
del
is C2H4 4.552 0.0305
use
ful because it makes it so simple to interpret the parameters in terms of molecular size and intermolecular forces. But it does have
limitations as well (as is the case of every scientific model!) Some other useful two-parameter and three-parameter (or more)
equations of state include the Redlich-Kwong, Dieterici, and Clausius models (Table 2.7.2). These have the advantage that they
allow for temperature dependence on some of the parameters, which as will be seen later, is necessary to model certain behaviors of
real gases.
Table 2.7.2 : Other equations of State
Model Equation of State
RT
Ideal p =
Vm
RT a
van der Waals (van der Waals J. D., 1967) p = −
2
Vm − b Vm
2.7.1 https://chem.libretexts.org/@go/page/84445
Model Equation of State
RT a
Redlich-Kwong (Redlich & Kwong, 1949) p = −
−
−
Vm − b √T Vm (Vm + b)
RT −a
Dieterici (Dieterici, 1899) p = exp( )
Vm − b Vm RT
RT a
Clausius p = −
2
Vm − b T (Vm + c)
RT B(T ) C(T )
Virial Equations p = (1 + + …)
Vm Vm Vm
RT B(T ) C (T )
p = (1 + + …) (2.7.3)
Vm Vm Vm
In the limit that B(T) (the Second Virial Coefficient) and C(T) are zero, the equation becomes the ideal gas law. Also, the molar
volume of gases are small, the contributions from the third, fourth, etc. terms decrease in magnitude, allowing one to truncate the
series at a convenient point. The second virial coefficient can be predicted from a theoretical intermolecular potential function by
∞
U (r)
2
B(T ) = Na ∫ [1 − exp( )] 2π r dr
r=0
kB T
The quality of an intermolecular potential can be determined (partially) by the potential’s ability to predict the value of the second
virial coefficient, B(T ).
∞ for r ≤ σ
U (r) = {
0 for r > σ.
In this function, σ is determined by the size of the molecules. If two molecules come within a distance r of one another, they
collide, bouncing off in a perfectly elastic collision. Real molecules, however, with have a range of intermolecular separations
through which they will experience attractive forces (the so-called “soft wall” of the potential surface.) And then at very small
separations, the repulsive forces will dominate, pushing the molecules apart (the so-called “hard wall” of the potential surface.)
A commonly used intermolecular potential, U (r) , is the Leonard-Jones potential. This function has the form
⎡ ⎤
σ 12 σ 6
⎢ ⎥
U (r) = 4ϵ ⎢ ( ) − ( ) ⎥
⎢ r r ⎥
⎣ ⎦
repulsive term attractive term
where σ governs the width of the potential well, and ϵ governs the depth. The distance between molecules is given by r. The
repulsive interactions between molecules are contained in the first terms and the attractive interactions are found in the second
term.
2.7.2 https://chem.libretexts.org/@go/page/84445
Figure 2.7.2 : The estimation of Lennard-Jones potential parameters for mixed pairs of atoms. (CC BY-SA 4.0; Cnrowley).
A commonly used method of creating a power series based on another equation is the Taylor Series Expansion. This is an
expansion of a function about a useful reference point where each of the terms is generated by differentiating the original
function.
For a function f (x), the Taylor series F (x) can be generated from the expression
2
d ∣ 1 d ∣
2
F (x) = f (a) + f (x)∣ (x − a) + f (x)∣ (x − a) +…
∣ 2
dx x=a
2! dx ∣
x=a
This can be applied to any equation of state to derive an expression for the virial coefficients in terms of the parameters of the
equation of state.
The van der Waals equation can be written in terms of molar volume (Equation 2.7.2 ). When multiplying the right hand side
by (where u = 1/v) yields:
u
RT u 2
p = − au
1 − bu
This expression can be "Talyor" expanded (to the first three terms) about u =0 (which corresponds to an infinite molar
volume.) The coefficient terms that are needed for the expansion are
p(u = 0) = 0
dp RT bRT u
∣
∣ = [ + − 2au] = RT
u=0
du 1 − bu (1 − bu)2
u=0
2 2
d p 1 bRT bRT 2 b RT u
∣
∣ = [ + − − 2au] = RT − a
2 u=0 2 2 3
du 2 (1 − bu) (1 − bu) (1 − bu) u=0
3
d p
2
∣
∣u=0 = RT b
3
du
And the virial equation can then be expressed in terms of the van der Waals parameters as
3
2 2(u)
p = 0 + RT (u) + (bRT − a)(u ) + RT b
2.7.3 https://chem.libretexts.org/@go/page/84445
And the second virial coefficient is given by
a
B(T ) = b −
RT
where V is the molar volume. For an ideal gas, Z = 1 under all combinations of P , V , and T . However, real gases will show
m m
some deviation (although all gases approach ideal behavior at low p, high Vm, and high T.) The compression factor for nitrogen at
several temperatures is shown below over a range of pressures.
or
⎛ ⎞
∂Z
lim ⎜ ⎟ =0
1/ Vm →0 1
⎝∂( )⎠
Vm
Using the virial equation of state (Equation 2.7.3 ), the Boyle temperature can be expressed in terms of the virial coefficients.
Starting with the compression factor
B
Z =1+ +…
Vm
2.7.4 https://chem.libretexts.org/@go/page/84445
Critical Behavior
The isotherms (lines of constant temperature) of CO2 reveal a very large deviation from ideal behavior. At high temperatures, CO2
behaves according to Boyle’s Law. However, at lower temperatures, the gas begins to condense to form a liquid at high pressures.
At one specific temperature, the critical temperature, the isotherm begins to display this critical behavior. The temperature,
pressure, and molar volume (p , T , and V ) at this point define the critical point. In order to solve for expressions for the critical
c c c
constants, one requires three equations. The equation of state provides one relationship. The second can be generated by
recognizing that the slope of the isotherm at the critical point is zero. And finally, the third expression is derived by recognizing
that the isotherm passes through an inflection point at the critical point. Using the van der Waals equation as an example, these
three equations can be generated as follows:
a
pc =
2
27b
8a
Tc =
27bR
Vc = 3b
The critical variables can be used in this fashion to determine the values of the molecular parameters used in an equation of state
(such as the van der Waals equation) for a given substance.
is very nearly the same for any substance. This is consistent with what is predicted by the van der Waals equation, which predicts
Z = 0.375 irrespective of substance.
c
V
Vr =
Vc
T
Tr =
Tc
several physical properties are found to be comparable for real substances. For example (Guggenheim, 1945), for argon, krypton,
nitrogen, oxygen, carbon dioxide and methane the reduced compressibility is
pc Vc
[ ≈ 0.292
RTc
2.7.5 https://chem.libretexts.org/@go/page/84445
Also, the reduced compression factor can be plotted as a function of reduced pressure for several substances at several reduced
isotherms with surprising consistency irrespective of the substance:
This page titled 2.7: Real Gases is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick Fleming.
2.7.6 https://chem.libretexts.org/@go/page/84445
2.E: Gases (Exercises)
Q2.1
Assuming the form of the Maxwell distribution allowing for motion in three directions to be
2
mv
2
f (v) = N v exp ( ) (2.E.1)
2 kB T
derive the correct expression for N such that the distribution is normalized. Hint: a table of definite integrals indicates
∞ −
2n −ax
2 1 √π
∫ x e dx = (2.E.2)
0
4 a3/2
Q2.2
Dry ice (solid CO2) has a density of 1.6 g/cm3. Assuming spherical molecules, estimate the collisional cross section for C O . How 2
Q2.3
Calculate the pressure exerted by 1.00 mol of Ar, N2, and CO2 as an ideal gas, a van der Waals gas, and a Redlich-Kwong gas, at
25 °C and 24.4 L.
Q2.4
The compression factor Z for CO2 at 0 °C and 100 atm is 0.2007. Calculate the volume of a 2.50 mole sample of CO2 at 0 °C and
100 atm.
Q2.5
Ar N2 CO2
ideal
Redlich-Kwong
Q2.6
What is the maximum pressure that will afford a N2 molecule a mean-free-path of at least 1.00 m at 25 °C?
Q2.7
In a Knudsen cell, the effusion orifice is measured to be 0.50 mm2. If a sample of naphthalene is allowed to effuse for 1.0 hr at a
temperature of 40.3 °C, the cell loses 0.0236 g. From this data, calculate the vapor pressure of naphthalene at this temperature.
Q2.8
The vapor pressure of scandium was determined using a Knudsen cell [Kirkorian, J. Phys. Chem., 67, 1586 (1963)]. The data from
the experiment are given below.
Temperature 1555.4 K
2.E.1 https://chem.libretexts.org/@go/page/84446
From this data, find the vapor pressure of scandium at 1555.4 K.
Q2.9
A thermalized sample of gas is one that has a distribution of molecular speeds given by the Maxwell-Boltzmann distribution.
Considering a sample of N2 at 25 cC what fraction of the molecules have a speed less than
a. the most probably speed
b. the average sped
c. the RMS speed?
d. The RMS speed of helium atoms under the same conditions?
Q2.10
Assume that a person has a body surface area of 2.0 m2. Calculate the number of collisions per second with the total surface area of
this person at 25 °C and 1.00 atm. (For convenience, assume air is 100% N2)
Q2.11
Two identical balloons are inflated to a volume of 1.00 L with a particular gas. After 12 hours, the volume of one balloon has
decreased by 0.200 L. In the same time, the volume of the other balloon has decreased by 0.0603 L. If the lighter of the two gases
was helium, what is the molar mass of the heavier gas?
Q2.12
Assuming it is a van der Waals gas, calculate the critical temperature, pressure and volume for C O . 2
Q2.13
Find an expression in terms of van der Waals coefficients for the Boyle temperature. (Hint: use the viral expansion of the van der
Waals equation to find an expression for the second viral coefficient!)
Q2.14
Consider a gas that follows the equation of state
RT
p = (2.E.3)
Vm − b
Using a virial expansion, find an expression for the second virial coefficient.
Q2.15
Consider a gas that obeys the equation of state
nRT
p = (2.E.4)
Vm − b
where a and b are non-zero constants. Does this gas exhibit critical behavior? If so, find expressions for p , V , and T in terms of
c c c
Q2.16
Consider a gas that obeys the equation of state
nRT an
p = − (2.E.5)
V − nB V
a. Find an expression for the Boyle temperature in terms of the constant a , b , and R .
b. Does this gas exhibit critical behavior? If so, find expressions for p , V , and T in terms of the constants a , b , and R .
c c c
This page titled 2.E: Gases (Exercises) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
2.E.2 https://chem.libretexts.org/@go/page/84446
2.S: Gases (Summary)
Learning Objectives
After mastering the material covered in this chapter, one will be able to:
1. Understand the relationships demonstrated by and perform calculations using the empirical gas laws (Boyle’s Law, Charles’
Law, Gay-Lussac’s Law, and Avogadro’s Law, as well as the combined gas law.)
2. Understand and be able to utilize the ideal gas law in applications important in chemistry.
3. State the postulates of the Kinetic Molecular theory of gases.
4. Utilize the Maxwell and Maxwell-Boltzmann distributions to describe the relationship between temperature and the distribution
of molecular speeds.
5. Derive an expression for pressure based on the predictions of the kinetic molecular theory for the collisions of gas molecules
with the walls of a container.
6. Derive and utilize an expression for the frequency with which molecules in a gas sample collide with other molecules.
7. Derive and utilize an expression for the mean-free-path of molecules based on temperature, pressure, and collisional cross
section.
8. Explain how the van der Waals (and other) model(s) allow for deviations from ideal behavior of gas samples.
9. Derive an expression for the Boyle temperature and interpret the results based on how a gas’s behavior approaches that of an
ideal gas.
10. Explain and utilize the Principle of Corresponding States.
2.S.1 https://chem.libretexts.org/@go/page/84447
van der Waals’ equation
Virial Equation
References
1. Avogadro, A. (1811). Essay on a Manner of Determining the Relative Masses of the Elementary Molecules of Bodies, and the
Proportions in Which They Enter into These Compounds. Journal de Physique, 73, 58-76.
2. Bernoulli, D. (1738). Hydronamica.
3. Clausius, R. (1857). Ueber die Art der Bewegung, welche wir Wärme nennen. Annalen der Physik, 176(3), 353–379.
doi:10.1002/andp.18571760302
4. Dieterici, C. (1899). Ann. Phys. Chem., 69, 685.
5. Einstein, A. (1905). Über die von der molekularkinetischen Theorie der Wärme geforderte Bewegung von in ruhenden
Flüssigkeiten suspendierten Teilchen. Annalen der Physik, 17(8), 549-560. doi:10.1002/andp.19053220806
6. Encycolopedia, N. W. (2016). Amedeao Avogadro. Retrieved April 13, 2016, from New World Encycolpedia:
http://www.newworldencyclopedia.org/...medeo_Avogadro
7. Fazio, F. (1992). Using Robert Boyle's Original Data in the Physics and Chemistry Classrooms. Journal of College Science
Teaching, 363-365.
8. Guggenheim, E. A. (1945). Corresponding State for Perfect Liquids. Journal of Chemical Physics, 13, 253-261.
9. Hunter, M. (2004). Robert Boyle (1627 - 91). Retrieved March 10, 2016, from The Robert Boyle Project:
http://www.bbk.ac.uk/boyle/
10. Johannes Diderik van der Waals - Biographical. (2014). Retrieved March 12, 2016, from Nobelprize.org:
http://www.nobelprize.org/nobel_priz...waals-bio.html
11. Maxwell, J. C. (1860). Illustrations of the dynamical theory of gases. Part 1. On the motions and collisions of perfectly elastic
spheres. Phil. Mag., XIX, 19-32.
12. Maxwell, J. C. (1873). Molecules. Nature, 417, 903-915. doi:10.1038/417903a
13. Redlich, O., & Kwong, J. N. (1949). On the Thermodynamics of Solutions. V. An Equation of State. Fugacities of Gaseous
Solutions. Chemical Reviews, 44(1), 233-244.
14. van der Waals, J. D. (1913). The law of corresponding states for different substances. Proceedings of the Koninklijke
Nederlandse Akademie van Wetenschappen, (pp. 971-981).
15. van der Waals, J. D. (1967). The equation of state for gases and liquids. Nobel Lectures in Physics 1901 - 1921, 254-265.
This page titled 2.S: Gases (Summary) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
2.S.2 https://chem.libretexts.org/@go/page/84447
CHAPTER OVERVIEW
This page titled 3: First Law of Thermodynamics is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
1
3.1: Prelude to Thermodynamics
Albert Einstein, a noted physicist, said of thermodynamics (Einstein, 1979)
“A law is more impressive the greater the simplicity of its premises, the more different are
the kinds of things it relates, and the more extended its range of applicability. (..) It is the
only physical theory of universal content, which I am convinced, that within the
framework of applicability of its basic concepts will never be overthrown.”
Thermodynamics is the study of how energy flows into and out of systems and how it flows through the universe. People have been
studying thermodynamics for a very long time and have developed the field a great deal, including the incorporation of high-level
mathematics into the process. Many of the relationships may look cumbersome or complicated, but they are always describing the
same basic thing: the flow of energy through the universe.
Energy, of course, can be used to do many useful things, such as allow us to drive our cars, use electronic devices, heat our homes,
and cook our food. Chemistry is important as well since many of the processes in which we generate energy depend on chemical
reactions (such as the combustion of hydrocarbons to generate heat or electron transfer reactions to generate electron flow.) The
previous chapter investigated gases which are convenient systems to use to frame many discussions of thermodynamics since they
can be modeled using specific equations of state such as the ideal gas law or the van der Waals law. These relationships depend on
an important class of variables known as state variables.
State variables are those variables which depend only upon the current conditions affecting a system. Pressure, temperature and
molar volume are examples of state variables. A number of variables required to describe the flow of energy in a system do depend
on the pathway a system follows to come into its current state.
To illustrate the difference, consider climbing a mountain. You may choose to walk straight up the side of the mountain, or you
may choose to circle the mountain several times in order to get to the top. These two pathways will differ in terms of how far you
actually walk (a path-dependent variable) to attain the same change in altitude (an example of a state variable.)
This page titled 3.1: Prelude to Thermodynamics is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
3.1.1 https://chem.libretexts.org/@go/page/84302
3.2: Work and Heat
One of the pioneers in the field of modern thermodynamics was James P. Joule (1818 - 1889). Among the experiments Joule
carried out, was an attempt to measure the effect on the temperature of a sample of water that was caused by doing work on the
water. Using a clever apparatus to perform work on water by using a falling weight to turn paddles within an insulated canister
filled with water, Joule was able to measure a temperature increase in the water.
Figure 3.2.1 : (left) James Prescott Joule (1818 - 1889) (right) Joule's apparatus for measuring the work equivalent of heat. (CC BY-
SA 3.0; Dr. Mirko Junge)
Thus, Joule was able to show that work and heat can have the same effect on matter – a change in temperature! It would then be
reasonable to conclude that heating, as well as doing work on a system will increase its energy content, and thus it’s ability to
perform work in the surroundings. This leads to an important construct of the First Law of Thermodynamics:
The capacity of a system to do work is increased by heating the system or doing work on
it.
The internal energy (U) of a system is a measure of its capacity to supply energy that can do work within the surroundings,
making U the ideal variable to keep track of the flow of heat and work energy into and out of a system. Changes in the internal
energy of a system (ΔU ) can be calculated by
ΔU = Uf − Ui (3.2.1)
where the subscripts i and f indicate initial and final states of the system. U as it turns out, is a state variable. In other words, the
amount of energy available in a system to be supplied to the surroundings is independent on how that energy came to be available.
That’s important because the manner in which energy is transferred is path dependent.
There are two main methods energy can be transferred to or from a system. These are suggested in the previous statement of the
first law of thermodynamics. Mathematically, we can restate the first law as
ΔU = q + w
or
dU = dq + dw
where q is defined as the amount of energy that flows into a system in the form of heat and w is the amount of energy lost due to
the system doing work on the surroundings.
Heat
Heat is the kind of energy that in the absence of other changes would have the effect of changing the temperature of the system. A
process in which heat flows into a system is endothermic from the standpoint of the system (q > 0, q
system < 0 ).
surroundings
Likewise, a process in which heat flows out of the system (into the surroundings) is called exothermic (q < 0, system
qsurroundings> 0 ). In the absence of any energy flow in the form or work, the flow of heat into or out of a system can be measured
by a change in temperature. In cases where it is difficult to measure temperature changes of the system directly, the amount of heat
3.2.1 https://chem.libretexts.org/@go/page/84303
energy transferred in a process can be measured using a change in temperature of the soundings. (This concept will be used later in
the discussion of calorimetry).
An infinitesimal amount of heat flow into or out of a system can be related to a change in temperature by
dq = C dT
Heat capacities generally have units of (J mol-1 K-1) and magnitudes equal to the number of J needed to raise the temperature of 1
mol of substance by 1 K. Similar to a heat capacity is a specific heat which is defined per unit mass rather than per mol. The
specific heat of water, for example, has a value of 4.184 J g-1 K-1 (at constant pressure – a pathway distinction that will be
discussed later.)
How much energy is needed to raise the temperature of 5.0 g of water from 21.0 °C to 25.0 °C?
Solution
q = mC ΔT
J
= (5.0 g )(4.184 )(25.0 °C − 21.0 °C )
g °C
= 84 J
A partial derivative, like a total derivative, is a slope. It gives a magnitude as to how quickly a function changes value when
one of the dependent variables changes. Mathematically, a partial derivative is defined for a function f (x , x , … x ) by
1 2 n
Because it measures how much a function changes for a change in a given dependent variable, infinitesimal changes in the in
the function can be described by
∂f
df = ∑ ( )
∂xi
i xj ≠i
So that each contribution to the total change in the function f can be considered separately.
For simplicity, consider an ideal gas. The pressure can be calculated for the gas using the ideal gas law. In this expression,
pressure is a function of temperature and molar volume.
RT
p(V , T ) =
V
and
∂p R
( ) = (3.2.3)
∂T V
V
3.2.2 https://chem.libretexts.org/@go/page/84303
So that the change in pressure can be expressed
∂p ∂p
dp = ( ) dV + ( ) dT (3.2.4)
∂V ∂T
T V
Macroscopic changes can be expressed by integrating the individual pieces of Equation 3.2.4 over appropriate intervals.
V2 T2
∂p ∂p
Δp = ∫ ( ) dV + ∫ ( ) dT
V1
∂V T1
∂T
T V
This can be thought of as two consecutive changes. The first is an isothermal (constant temperature) expansion from V to V 1 2
at T and the second is an isochoric (constant volume) temperature change from T to T at V . For example, suppose one
1 1 2 2
needs to calculate the change in pressure for an ideal gas expanding from 1.0 L/mol at 200 K to 3.0 L/mol at 400 K. The set up
might look as follows.
V2 T2
RT R
Δp = ∫ (− ) dV + ∫ ( ) dT
2
V1 V T1
V
isothermal expansion isochoric heating
or
3.0 L/mol 400, K
R(400 K) R
Δp = ∫ (− ) dV + ∫ ( ) dT
2
1.0 L/mol V 200 K 1.0 L/mol
= −5.47 atm
Alternatively, one could calculate the change as an isochoric temperature change from T1 to T2 at V1 followed by an
isothermal expansion from V to V at T :
1 2 2
T2 V2
R RT
Δp = ∫ ( ) dT + ∫ (− ) dV
2
T1
V V1 V
or
3.2.3 https://chem.libretexts.org/@go/page/84303
400, K 3.0 L/mol
R R(400 K)
Δp = ∫ ( ) dT + ∫ (− ) dV
2
200 K 1.0 L/mol 1.0 L/mol V
= −5.47 atm
This results demonstrates an important property of pressure in that pressure is a state variable, and so the calculation of
changes in pressure do not depend on the pathway!
Work
Work can take several forms, such as expansion against a resisting pressure, extending length against a resisting tension (like
stretching a rubber band), stretching a surface against a surface tension (like stretching a balloon as it inflates) or pushing electrons
through a circuit against a resistance. The key to defining the work that flows in a process is to start with an infinitesimal amount of
work defined by what is changing in the system.
Table 3.1.1: Changes to the System
Type of work Displacement Resistance dw
The pattern followed is always an infinitesimal displacement multiplied by a resisting force. The total work can then be determined
by integrating along the pathway the change follows.
What is the work done by 1.00 mol an ideal gas expanding from a volume of 22.4 L to a volume of 44.8 L against a constant
external pressure of 0.500 atm?
Solution
dw = −pext dV
since the pressure is constant, we can integrate easily to get total work
V2
w = −pexp ∫ dV
V1
= −pexp (V2 − V1 )
8.314 J
= −(0.500 am)(44.8 L − 22.4 L) ( )
0.08206 atm L
= −1130 J = −1.14 kJ
Note: The ratio of gas law constants can be used to convert between atm·L and J quite conveniently!
This page titled 3.2: Work and Heat is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick Fleming.
3.2.4 https://chem.libretexts.org/@go/page/84303
3.3: Reversible and Irreversible Pathways
The most common example of work in the systems discussed in this book is the work of expansion. It is also convenient to use the
work of expansion to exemplify the difference between work that is done reversibly and that which is done irreversibly. The
example of expansion against a constant external pressure is an example of an irreversible pathway. It does not mean that the gas
cannot be re-compressed. It does, however, mean that there is a definite direction of spontaneous change at all points along the
expansion.
Imagine instead a case where the expansion has no spontaneous direction of change as there is no net force push the gas to seek a
larger or smaller volume. The only way this is possible is if the pressure of the expanding gas is the same as the external pressure
resisting the expansion at all points along the expansion. With no net force pushing the change in one direction or the other, the
change is said to be reversible or to occur reversibly. The work of a reversible expansion of an ideal gas is fairly easy to calculate.
dw = −pdV
or
w = −∫ pdV
But now that the external pressure is not constant, p cannot be extracted from the integral. Fortunately, however, there is a simple
relationship that tells us how p changes with changing V – the equation of state! If the gas is assumed to be an ideal gas
nRT
w = −∫ pdV − ∫ ( ) dV
V
And if the temperature is held constant (so that the expansion follows an isothermal pathway) the nRT term can be extracted from
the integral.
V2
dV V1
w = −nRT ∫ = −nRT ln( ) (3.3.1)
V1
V V2
Equation 3.3.1 is derived for ideal gases only; a van der Waal gas would result in a different version.
What is the work done by 1.00 mol an ideal gas expanding reversibly from a volume of 22.4 L to a volume of 44.8 L at a
constant temperature of 273 K?
Solution
Using Equation 3.3.1 to calculate this
J 44.8 L
w = −(1.00 mol ) (8.314 ) (273 K ) ln( )
mol K 22.4 L
= −1570 J = 1.57 kJ
Note: A reversible expansion will always require more work than an irreversible expansion (such as an expansion against a
constant external pressure) when the final states of the two expansions are the same!
3.3.1 https://chem.libretexts.org/@go/page/84304
The work of expansion can be depicted graphically as the area under the p-V curve depicting the expansion. Comparing examples
3.3.1 and 3.1.2, for which the initial and final volumes were the same, and the constant external pressure of the irreversible
expansion was the same as the final pressure of the reversible expansion, such a graph looks as follows.
The work is depicted as the shaded portion of the graph. It is clear to see that the reversible expansion (the work for which is
shaded in both light and dark gray) exceeds that of the irreversible expansion (shaded in dark gray only) due to the changing
pressure of the reversible expansion. In general, it will always be the case that the work generated by a reversible pathway
connecting initial and final states will be the maximum work possible for the expansion.
It should be noted (although it will be proven in a later chapter) that ΔU for an isothermal reversible process involving only p-V
work is 0 for an ideal gas. This is true because the internal energy, U, is a measure of a system’s capacity to convert energy into
work. In order to do this, the system must somehow store that energy. The only mode in which an ideal gas can store this energy is
in the translational kinetic energy of the molecules (otherwise, molecular collisions would not need to be elastic, which as you
recall, was a postulate of the kinetic molecular theory!) And since the average kinetic energy is a function only of the temperature,
it (and therefore U ) can only change if there is a change in temperature. Hence, for any isothermal process for an ideal gas,
ΔU = 0 . And, perhaps just as usefully, for an isothermal process involving an ideal gas, q = −w , as any energy that is expended
by doing work must be replaced with heat, lest the system temperature drop.
However, dV = 0 since the volume is constant! As such, dU can be expressed only in terms of the heat that flows into or out of
the system at constant volume
dU = dqv
This suggests an important definition for the constant volume heat capacity (C ) which is
V
∂U
CV ≡ ( )
∂T
V
3.3.2 https://chem.libretexts.org/@go/page/84304
When Equation 3.3.2 is integrated the
T2
q =∫ nCV dt (3.3.3)
T1
Consider 1.00 mol of an ideal gas with C = 3/2R that undergoes a temperature change from 125 K to 255 K at a constant
V
Solution
Since this is a constant volume process
w =0
q =∫ nCV dt
T1
T2
q = nCV ∫ dt
T1
= nCV (T2 − T1 )
3 J
= (1.00 mol) ( 8.314 ) (255 K − 125 K)
2 mol K
= 1620 J = 1.62 kJ
= 1.62 kJ
H ≡ U + pV
or
dH ≡ dU + d(pV )
= dU + pdV + V dp
For reversible changes at constant pressure (dp = 0 ) for which only p-V work is done
dH = dq + dw + pdV + V dp (3.3.4)
= dq (3.3.6)
And just as in the case of constant volume changes, this implies an important definition for the constant pressure heat capacity
3.3.3 https://chem.libretexts.org/@go/page/84304
∂H
Cp ≡ ( )
∂T
p
Solution
T2
q =∫ nCp dT
T1
T2
q = nCp ∫ dT
T1
= nCp (T2 − T1 )
5 J
= (1.00 mol) ( 8.314 ) (255 K − 125 K) = 2700 J = 1.62 kJ
2 mol K
So via Equation 3.3.6 (specifically the integrated version of it using differences instead of differentials)
ΔH = q = 1.62 kJ
ΔU = ΔH − Δ(pV )
= ΔH − nRΔT
J
= 2700 J − (1.00 mol) (8.314 ) (255 K − 125 K)
mol K
= 1620 J = 1.62 kJ
Calculate q, w, ΔU , and ΔH for 1.00 mol of an ideal gas expanding reversibly and isothermally at 273 K from a volume of
22.4 L and a pressure of 1.00 atm to a volume of 44.8 L and a pressure of 0.500 atm.
Solution
Since this is an isothermal expansion, Equation3.3.1 is applicable
3.3.4 https://chem.libretexts.org/@go/page/84304
V2
w = −nRT ln
V1
J 44.8 L
= (1.00 mol) (8.314 ) (255 K) ln( )
mol K 22.4 L
= 1572 J = 1.57 kJ
ΔU = q + w
= q + 1.57 KJ
=0
q = −1.57 kJ
ΔH = ΔU + Δ(pV ) = 0 + 0
Adiabatic Pathways
An adiabatic pathway is defined as one in which no heat is transferred (q = 0 ). Under these circumstances, if an ideal gas expands,
it is doing work (w < 0 ) against the surroundings (provided the external pressure is not zero!) and as such the internal energy must
drop (ΔU < 0 ). And since ΔU is negative, there must also be a decrease in the temperature (ΔT < 0 ). How big will the decrease
in temperature be and on what will it depend? The key to answering these questions comes in the solution to how we calculate the
work done.
If the adiabatic expansion is reversible and done on an ideal gas,
dw = −pdV
and
dw = nCv dT (3.3.7)
−pdV = nCv dT
And rearranging to gather the temperature terms on the right and volume terms on the left yields
dV CV dT
=−
V R T
This expression can be integrated on the left between V and V and on the right between T and T . Assuming that
1 2 1 2 Cv /nR is
independent of temperature over the range of integration, it can be pulled from the integrand in the term on the right.
V2 T2
dV CV dT
∫ =− ∫
V1
V R T1
T
The result is
V2 CV T2
ln( ) =− ln( )
V1 R T1
or
C
V
−
V2 T2 R
( ) =( )
V1 T1
3.3.5 https://chem.libretexts.org/@go/page/84304
or
C C
V V
R R
V1 T = V2 T
1 2
or
R
−
V1 C
V
T1 ( ) = T2 (3.3.8)
V2
Example 3.3.5:
1.00 mol of an ideal gas (CV = 3/2 R) initially occupies 22.4 L at 273 K. The gas expands adiabatically and reversibly to a final
volume of 44.8 L. Calculate ΔT , q, w, ΔU , and ΔH for the expansion.
Solution
Since the pathway is adiabatic:
q =0
T2 = T1 ( )
V2
2/3
22.4 L
= (273 K)( )
44.8 L
= 172 K
So
w = ΔU = nCv ΔT
3 J
= (1.00 mol) ( 8.314 ) (−101 K)
2 mol K
= 1.260 kJ
ΔH = ΔU + nRΔT
3 J
= −1260 J + (1.00 mol) ( 8.314 ) (−101 K)
2 mol K
= −2100 J
The following table shows recipes for calculating q, w , ΔU , and ΔH for an ideal gas undergoing a reversible change along the
specified pathway.
Table 3.2.1: Thermodynamics Properties for a Reversible Expansion or Compression
Pathway q w ΔU ΔH
Isochoric CV ΔT 0 CV ΔT CV ΔT + V Δp
adiabatic 0 CV ΔT CV ΔT Cp ΔT
3.3.6 https://chem.libretexts.org/@go/page/84304
This page titled 3.3: Reversible and Irreversible Pathways is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated
by Patrick Fleming.
3.3.7 https://chem.libretexts.org/@go/page/84304
3.4: Calorimetry
As chemists, we are concerned with chemical changes and reactions. The thermodynamics of chemical reactions can be very
important in terms of controlling the production of desired products and preventing safety hazards such as explosions. As such,
measuring and understanding the thermochemistry of chemical reactions is not only useful, but essential!
Calorimetry
The techniques of calorimetry can be used to measure q for a chemical reaction directly. The enthalpy change for a chemical
reaction is of significant interest to chemists. An exothermic reaction will release heat (q < 0, q
reaction > 0 ) causing the
surroundings
from the surroundings, causing the temperature of the surrounding to drop. Measuring the temperature change in the surroundings
allows for the determination of how much heat was released or absorbed in the reaction.
Bomb Calorimetry
Bomb calorimetry is used predominantly to measure the heat evolved in combustion reactions, but can be used for a wide variety of
reactions. A typical bomb calorimetry set up is shown here. The reaction is contained in a heavy metallic container (the bomb)
forcing the reaction to occur at constant volume. As such, the heat evolved (or absorbed) by the reaction is equal to the change in
internal energy (DUrxn). The bomb is then submerged in a reproducible quantity of water, the temperature of which is monitored
with a high-precision thermometer.
For combustion reactions, the bomb will be loaded with a small sample of the compound to be combusted, and then the bomb is
filled with a high pressure (typically about 10 atm) of O2. The reaction is initiated by supplying heat using a short piece of resistive
wire carrying an electrical current.
3.4.1 https://chem.libretexts.org/@go/page/84305
Figure 3.4.1 : A Bomb Calorimeter. After the temperature of the water in the insulated container has reached a constant value, the
combustion reaction is initiated by passing an electric current through a wire embedded in the sample. Because this calorimeter
operates at constant volume, the heat released is not precisely the same as the enthalpy change for the reaction. (CC BY-SA-NC;
Anonymous by request).
The calorimeter must be calibrated by carrying out a reaction for which ΔU is well known, so that the resulting temperature
rxn
change can be related to the amount of heat released or absorbed. A commonly used reaction is the combustion of benzoic acid.
This makes a good choice since benzoic acid reacts reliably and reproducibly under normal bomb calorimetry conditions. The
“water equivalent” of the calorimeter can then be calculated from the temperature change using the following relationship:
nΔUc + ewrire + eother
W =
ΔT
where n is the number of moles of benzoic acid used, ΔU is the internal energy of combustion for benzoic acid (3225.7 kJ mol-1 at
c
25 oC), ewire accounts for the energy released in the combustion of the fuse wire, eother account for any other corrections (such as
heat released due to the combustion of residual nitrogen in the bomb), and DT is the measured temperature change in the
surrounding water bath.
Once the “water equivalent” is determined for a calorimeter, the temperature change can be used to find ΔUc for an unknown
compound from the temperature change created upon combustion of a known quantity of the substance.
W ΔT − ewire − eother
ΔUc =
nsample
The experiment above is known as “isothermal bomb calorimetry” as the entire assembly sits in a constant temperature laboratory.
Another approach is to employ “adiabatic bomb calorimetry” in which the assembly sits inside of a water jacket, the temperature of
which is controlled to match the temperature of the water inside the insulated container. By matching this temperature, there is no
thermal gradient, and thus no heat leaks into or out of the assembly during an experiment (and hence the experiment is effectively
“adiabatic”).
Finding ΔU c
The enthalpy of combustion can be calculated from the internal energy change if the balanced chemical reaction is known. Recall
from the definition of enthalpy
ΔH = ΔU + Δ(pV )
and if the gas-phase reactants and products can be treated as ideal gases (pV = nRT )
3.4.2 https://chem.libretexts.org/@go/page/84305
ΔH = ΔU + RT Δngas
it can be seen that Δn gas is -0.5 mol of gas for every mole of benzoic acid reacted.
A student burned a 0.7842 g sample of benzoic acid (C H O ) in a bomb calorimeter initially at 25.0 oC and saw a
7 6 2
temperature increase of 2.02 oC. She then burned a 0.5348 g sample of naphthalene (C H ) (again from an initial temperature
10 8
of 25 oC) and saw a temperature increase of 2.24 oC. From this data, calculate ΔH for naphthalene (assuming ewire and eother
c
are unimportant.)
Solution
First, the water equivalent:
1 mol
[(0.7841 g) ( )] (3225.7 kJ/mol)
122.124 g
W = = 10.254 kJ/°C
2.02 °C
with Δn gas = −2 .
So
8.314
ΔHc = 5546.4 kJ/mol + ( kJ/(mol K)) (298 L)(−2) = 5541 kJ/mol
1000
The literature value (Balcan, Arzik, & Altunata, 1996) is 5150.09 kJ/mol. So that’s not too far off!
This page titled 3.4: Calorimetry is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick Fleming.
3.4.3 https://chem.libretexts.org/@go/page/84305
3.5: Temperature Dependence of Enthalpy
It is often required to know thermodynamic functions (such as enthalpy) at temperatures other than those available from tabulated
data. Fortunately, the conversion to other temperatures is not difficult.
At constant pressure
dH = Cp dT
T2
ΔH = ∫ Cp dT (3.5.1)
T2
ΔH = Cp ΔT (3.5.2)
If the temperature dependence of the heat capacity is known, it can be incorporated into the integral in Equation 3.5.1. A common
empirical model used to fit heat capacities over broad temperature ranges is
c
Cp (T ) = a + bT + (3.5.3)
2
T
After combining Equations 3.5.3 and 3.5.1, the enthalpy change for the temperature change can be found obtained by a simple
integration
T2
c
ΔH = ∫ (a + bT + ) dT (3.5.4)
2
T1 T
b 1 1
2 2
= a(T2 − T1 ) + (T −T )−c ( − ) (3.5.6)
2 1
2 T2 T1
This expression can then be used with experimentally determined values of a , b , and c , some of which are shown in the following
table.
Table 3.5.1 : Empirical Parameters for the temperature dependence of C p
H2O(l) 75.29 0 0
What is the molar enthalpy change for a temperature increase from 273 K to 353 K for Pb(s)?
Solution
The enthalpy change is given by Equation 3.5.1 with a temperature dependence Cp given by Equation 3.5.1 using the
parameters in Table 3.5.1. This results in the integral form (Equation 3.5.6):
3.5.1 https://chem.libretexts.org/@go/page/84306
b 1 1
2 2
ΔH = a(T2 − T1 ) + (T −T )−c ( − )
2 1
2 T2 T1
when substituted with the relevant parameters of Pb(s) from Table 3.5.1.
J
ΔH = (22.14 (353 K − 273 K)
mol K
−2 J
1.172 × 10 2
mol K 2 2
+ ((353 K ) − (273 K ) )
2
J K 1 1
4
− 9.6 × 10 ( − )
mol (353 K) (273 K)
J J J
ΔH = 1770.4 + 295.5 + 470.5
mol mol mol
J
=2534.4
mol
For chemical reactions, the reaction enthalpy at differing temperatures can be calculated from
T2
The enthalpy of formation of NH3(g) is -46.11 kJ/mol at 25 oC. Calculate the enthalpy of formation at 100 oC.
Solution
N (g) + 3 H (g) ⇌ 2 NH (g)
2 2 3
N2(g) 29.12
H2(g) 28.82
NH3(g) 35.06
J J J J
= −46110 + [2 (35.06 ) − (29.12 ) − 3 (28.82 )] (373 K − 298 K)
mol mol K mol K mol K
kJ
= −49.5
mol
This page titled 3.5: Temperature Dependence of Enthalpy is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by Patrick Fleming.
3.5.2 https://chem.libretexts.org/@go/page/84306
3.6: Reaction Enthalpies
Reaction enthalpies are important, but difficult to tabulate. However, because enthalpy is a state function, it is possible to use Hess’
Law to simplify the tabulation of reaction enthalpies. Hess’ Law is based on the addition of reactions. By knowing the reaction
enthalpy for constituent reactions, the enthalpy of a reaction that can be expressed as the sum of the constituent reactions can be
calculated. The key lies in the canceling of reactants and products that °Ccur in the “data” reactions but not in the “target reaction.
Example 3.6.1:
Find ΔH rxn for the reaction
Given
with ΔH 1 = −110.53 kJ
with ΔH 2 = −393.51 kJ
Solution
The target reaction can be generated from the data reactions.
plus
equals
so
2×ΔH1 = −787.02 kJ
2×ΔH2 = 221.06 kJ
Using this definition, a convenient reaction for which enthalpies can be measured and tabulated is the standard formation
reaction. This is a reaction which forms one mole of the substance of interest in its standard state from elements in their standard
3.6.1 https://chem.libretexts.org/@go/page/84307
states. The enthalpy of a standard formation reaction is the standard enthalpy of formation (ΔH ). Some examples are
f
o
N aC l(s) :
with ΔH o
f
= −411.2 kJ/mol
C3 H8 (g) :
with ΔH o
f
= −103.8 kJ/mol
It is important to note that the standard state of a substance is temperature dependent. For example, the standard state of water at
-10 °C is solid, whereas the standard state at room temperature is liquid. Once these values are tabulated, calculating reaction
enthalpies becomes a snap. Consider the heat combustion (ΔH ) of methane (at 25 °C) as an example.
c
The reaction can expressed as a sum of a combination of the following standard formation reactions.
with ΔH f
o
= −74.6 kJ/mol
with ΔH f
o
= −393.5 kJ/mol
with ΔH f
o
= −285.8 kJ/mol
The target reaction can be generated from the following combination of reactions
with ΔH f
o
= −1× [−74.6 kJ/mol] = 74.6 kJ/mol
with ΔH f
o
= −393.5 kJ/mol
with ΔH f
o
= 2× [−285.8 kJ/mol] = −571.6 kJ/mol .
with ΔH o
c = −890.5 kJ/mol
Alternately, the reaction enthalpy could be calculated from the following relationship
o o
ΔHrxn = ∑ ν ⋅ ΔH − ∑ ν ⋅ ΔH
f f
products reactants
where ν is the stoichiometric coefficient of a species in the balanced chemical reaction. For the combustion of methane, this
calculation is
3.6.2 https://chem.libretexts.org/@go/page/84307
o o o
Δrxn = (1 mol) (ΔH (C O2 )) + (2 mol) (ΔH (H2 O)) − (1 mol) (ΔH (C H4 )) (3.6.1)
f f f
A note about units is in order. Note that reaction enthalpies have units of kJ, whereas enthalpies of formation have units of kJ/mol.
The reason for the difference is that enthalpies of formation (or for that matter enthalpies of combustion, sublimation, vaporization,
fusion, etc.) refer to specific substances and/or specific processes involving those substances. As such, the total enthalpy change is
scaled by the amount of substance used. General reactions, on the other hand, have to be interpreted in a very specific way. When
examining a reaction like the combustion of methane
with ΔH rxn= −890.5 kJ . The correct interpretation is that the reaction of one mole of CH4(g) with two moles of O2(g) to form
one mole of CO2(g) and two moles of H2O(l) releases 890.5 kJ at 25 °C.
Ionization Reactions
Ionized species appear throughout chemistry. The energy changes involved in the formation of ions can be measured and tabulated
for several substances. In the case of the formation of positive ions, the enthalpy change to remove a single electron at 0 K is
defined as the ionization potential.
+ −
M (g) → M (g) + e
with ΔH (0K) ≡ 1 st
ionization potential (IP)
The removal of subsequent electrons requires energies called the 2nd Ionization potential, 3rd ionization potential, and so on.
+ 2+ −
M (g) → M (g) + e
with ΔH (0K) ≡ 2 nd
IP
2+ 3+ −
M (g) → M (g) + e
with ΔH (0K) ≡ 3 rd
IP
An atom can have as many ionization potentials as it has electrons, although since very highly charged ions are rare, only the first
few are important for most atoms.
Similarly, the electron affinity can be defined for the formation of negative ions. In this case, the first electron affinity is defined
by
− −
X(g) + e → X (g)
The minus sign is included in the definition in order to make electron affinities mostly positive. Some atoms (such as noble gases)
will have negative electron affinities since the formation of a negative ion is very unfavorable for these species. Just as in the case
of ionization potentials, an atom can have several electron affinities.
− − 2−
X (g) + e → X (g)
3.6.3 https://chem.libretexts.org/@go/page/84307
XY (g) → X(g) + Y (g)
with ΔH ≡ D(X − Y )
In this process, one adds energy to the reaction to break bonds, and extracts energy for the bonds that are formed.
In this reaction, five C-H bonds, one C-C bond, and one C-O bond, and one O=O bond must be broken. Also, four C=O bonds, and
one O-H bond are formed.
C-H 413
C-C 348
C-O 358
O=O 495
C=O 799
O-H 463
Because the bond energies are defined for gas-phase reactants and products, this method does not account for the enthalpy change
of condensation to form liquids or solids, and so the result may be off systematically due to these differences. Also, since the bond
enthalpies are averaged over a large number of molecules containing the particular type of bond, the results may deviate due to the
variance in the actual bond enthalpy in the specific molecule under consideration. Typically, reaction enthalpies derived by this
method are only reliable to within ± 5-10%.
This page titled 3.6: Reaction Enthalpies is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
3.6.4 https://chem.libretexts.org/@go/page/84307
3.7: Lattice Energy and the Born-Haber Cycle
An important enthalpy change is the Lattice Energy, which is the energy required to take one mole of a crystalline solid to ions in
the gas phase. For NaCl(s), the lattice energy is defined as the enthalpy of the reaction
+ −
NaCl(s) → Na (g) + Cl (g)
with ΔH f
(N aC l) .
The other pathway involves a series of steps that take the elements from neutral species in their standard states to ions in the gas
phase.
N a(s) → N a(g)
with ΔH sub (N a)
+ −
N a(g) → N a (g) + e
with 1 st
I P (N a)
½C l2 (g) → C l(g)
with ½D(C l − C l)
− −
C l(g) + e → Cl (g)
with 1 st
EA(C l)
+ −
Na (g) + C l (g) → N aC l(s)
with ΔH Lat (N aC l)
It should be clear that when added (after proper manipulation if needed), the second set of reactions yield the first reaction. Because
of this, the total enthalpy changes must all add.
st st
ΔHsub (N a) + 1 I P (N a) + ½D(C l − C l) + 1 EA(C l) + ΔHlat (N aC l) = ΔHf (N aC l)
This can be depicted graphically, the advantage being that arrows can be used to indicate endothermic or exothermic changes. An
example of the Born-Haber Cycle for NaCl is shown below.
3.7.1 https://chem.libretexts.org/@go/page/84456
Figure 3.6.1: the Born-Haber Cycle for NaCl.
In many applications, all but one leg of the cycle is known, and the job is to determine the magnitude of the missing leg.
K(s) → K(g)
Br (l) → Br (g)
2 2
Br (g) → 2 Br(g)
2
with 1 st
I P (K) = 419kJ/mol
−
Br(g) + e → Br−(g)
with 1 st
EA(Br) = 194kJ/mol
+
K (g) + Br−(g) → KBr(s)
Answer
ΔHf = −246 kJ/mol
Note: This cycle required the extra leg of the vaporization of Br2. Many cycles involve ions with greater than unit charge
and may require extra ionization steps as well!
This page titled 3.7: Lattice Energy and the Born-Haber Cycle is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by Patrick Fleming.
3.7.2 https://chem.libretexts.org/@go/page/84456
3.E: First Law of Thermodynamics (Exercises)
Q3.1
In the attempt to measure the heat equivalent of mechanical work (as Joule did in his famous experiment) a student uses an
apparatus similar to that shown below:
The 1.50 kg weight is lifted 30.0 cm against the force due to gravity (9.8 N). If the specific heat of water is 4.184 J/(g °C), what is
the expected temperature increase of the 1.5 kg of water in the canister?
Q3.2
1.00 mol of an ideal gas, initially occupying 12.2 L at 298 K, expands isothermally against a constant external pressure of 1.00 atm
until the pressure of the gas is equal to the external pressure. Calculate Δp, q, w, ΔU , and ΔH for the expansion.
Q3.3
Consider 1.00 mol of an ideal gas expanding isothermally at 298 K from an initial volume of 12.2 L to a final volume of 22.4 L.
Calculate Δp, q, w, ΔU , and ΔH for the expansion.
Q3.4
Consider 1.00 mol of an ideal gas (CV = 3/2 R) Occupying 22.4 L that undergoes an isochoric (constant volume) temperature
increase from 298 K to 342 K. Calculate Δp, q, w, ΔU , and ΔH for the change.
Q3.5
Consider 1.00 mol of an ideal gas (Cp = 5/2 R) initially at 1.00 atm that undergoes an isobaric expansion from 12.2 L to 22.4 L.
Calculate ΔT , q, w, ΔU , and ΔH for the change.
Q3.6
Consider 1.00 mol of an ideal gas (CV = 3/2 R) initially at 12.2 L that undergoes an adiabatic expansion to 22.4 L. Calculate ΔT , q,
w , ΔU , and ΔH for the change.
Q3.7
Derive an expression for the work of an isothermal, reversible expansion of a gas that follows the equation of state (in which a is a
parameter of the gas)
2
an
pV = nRT − (3.E.1)
V
from V to V .
1 2
Q3.8
Use the following data [Huff, Squitieri, and Snyder, J. Am. Chem. Soc., 70, 3380 (1948)] to calculate the standard enthalpy of
formation of tungsten carbide, W C (s).
3.E.1 https://chem.libretexts.org/@go/page/84457
Reaction ΔH
o
(kJ)
Q3.9
The standard molar enthalpy of combustion (ΔH ) of propane gas is given by
c
The standard molar enthalpy of vaporization (ΔH vap ) for liquid propane
C3 H8 (l) → C3 H8 (g) (3.E.3)
c. Calculate the standard internal energy change of combustion (ΔH ) for liquid propane. c
Q3.10
The enthalpy of combustion (ΔH ) of aluminum borohydride, Al(BH ) (l), was measured to be -4138.4 kJ/mol [Rulon and
c 4 3
Mason, J. Am. Chem. Soc., 73, 5491 (1951)]. The combustion reaction for this compound is given by
Given the following additional data, calculate the enthalpy of formation of Al(BH 4 )3 (g) .
: ΔH = −1669.8 kJ/mol
Al2 O3 (s) f
Al(BH ) (l): ΔH
4 3 = 30.125 kJ/mol
vap
Q3.11
The standard enthalpy of formation (ΔH ) for water vapor is -241.82 kJ/mol at 25 °C. Use the data in the following table to
o
H2(g) 28.84
O2(g) 29.37
H2O(g) 33.58
Q3.12
ΔCp = (1.00 + 2.00 × 10
−3
T ) J/K and ΔH 298 = −5.00 kJ for a dimerization reaction
2A → A2 (3.E.5)
Q3.13
From the following data, determine the lattice energy of BaBr . 2
3.E.2 https://chem.libretexts.org/@go/page/84457
with ΔH sub = 129 kJ/mol
with 1st
I P (K) = 589.8 kJ/mol
+ 2+ −
Ca (g) → C a (g) + e (3.E.10)
with 2nd
I P (K) = 1145.4 kJ/mol
−
Br(g) + e → Br − (g) (3.E.11)
with 1st
EA(Br) = 194 kJ/mol
−
C a(s) + Br (l) → C aBr2 (s) (3.E.12)
2
Q3.15
Using average bond energies (Table T3) estimate the reaction enthalpy for the reaction
C2 H4 + H Br → C2 H5 Br (3.E.13)
This page titled 3.E: First Law of Thermodynamics (Exercises) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by Patrick Fleming.
3.E.3 https://chem.libretexts.org/@go/page/84457
3.S: First Law of Thermodynamics (Summary)
Learning Objectives
After mastering the material covered in this chapter, one will be able to:
1. Define the internal energy of a system as a measure of its capacity to do work on its surroundings.
2. Define work and heat and relate them to changes in the internal energy of a system.
3. Explain the difference between path dependent variables and path independent variables.
4. Define enthalpy in terms of internal energy, pressure, and volume.
5. Calculate First Law quantities such as q, w, ΔU and ΔH , for an ideal gas undergoing changes in temperature, pressure, and/or
volume along isothermal, isobaric, isochoric, or adiabatic pathways.
6. Perform calculations using data collected using calorimetry (at either constant pressure or constant volume).
7. Write a formation reaction (the reaction for which the standard enthalpy of formation is defined) for any compound.
8. Use enthalpies of formation to calculate reaction enthalpies.
9. Estimate reaction enthalpies from average bond dissociation enthalpies.
10. Define and utilize enthalpies for phase changes such as ΔH , ΔH , and ΔH to calculate the heat energy transferred in
f us sub vap
3.S.1 https://chem.libretexts.org/@go/page/84458
References
1. Balcan, M., Arzik, S., & Altunata, T. (1996). The determination of the heats of combustion and the resonance energies of some
substituted naphthalenes. Thermichimica Acta, 278, 49-56.
2. Einstein, A. (1979). Autobiographical Notes. A Centennial Edition. Open Court Publishing Company.
3. Encyclopedia.com. (2008). James Prescott Joule. Retrieved March 30, 2016, from Complete Dictionary of Scientific
Biography: http://www.encyclopedia.com/doc/1G2-2830902225.html
Contributors
Patrick E. Fleming (Department of Chemistry and Biochemistry; California State University, East Bay)
This page titled 3.S: First Law of Thermodynamics (Summary) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by Patrick Fleming.
3.S.2 https://chem.libretexts.org/@go/page/84458
CHAPTER OVERVIEW
This page titled 4: Putting the First Law to Work is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
1
4.1: Prelude to Putting the First Law to Work
As has been seen in previous chapters, may important thermochemical quantities can be expressed in terms of partial derivatives.
Two important examples are the molar heat capacities C and C which can be expressed as
p V
∂H
Cp = ( )
∂T
p
and
∂U
CV = ( )
∂T
V
These are properties that can be measured experimentally and tabulated for many substances. These quantities can be used to
calculate changes in quantities since they represent the slope of a surface (H or U ) in the direction of the specified path (constant p
or V ). This allows us to use the following kinds of relationships:
∂H
dH = ( ) dT
∂T
p
and
∂H
ΔH = ∫ ( ) dT
∂T
p
Because thermodynamics is kind enough to deal in a number of state variables, the functions that define how those variable
change must behave according to some very well determined mathematics. This is the true power of thermodynamics!
This page titled 4.1: Prelude to Putting the First Law to Work is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by Patrick Fleming.
4.1.1 https://chem.libretexts.org/@go/page/91109
4.2: Total and Exact Differentials
The fact that we can define the constant volume heat capacity as
∂U
CV ≡ ( ) (4.2.1)
∂T
V
suggests that the internal energy depends very intimately on two variables: volume and temperature. In fact, we will see that for a
single component system, state variables are always determined when two state variables are defined. In the case of internal energy,
we might write U = f (V , T ) or U (V , T ) .
This suggests that the way to change U is to change either V or T (or both!) And if there is a mathematical function that relates the
internal energy to these two variables, it should easy to see how it changes when either (or both!) are changed. This can be written
as a total differential:
∂U ∂U
dU = ( ) dV + ( ) dT (4.2.2)
∂V ∂T
T V
Even without knowing the actually mathematical function relating the variables to the property, we can imagine how to calculate
changes in the property from this expression.
V2 T2
∂U ∂U
ΔU = ∫ ( ) dV + ∫ ( ) dT
V1 ∂V T T1 ∂T V
In words, this implies that we can think of a change in U occurring due to an isothermal change followed by an isochoric change.
And all we need to know is the slope of the surface in each pathway direction. There are a couple of very important experiments
people have done to explore the measurement of those kinds of slopes. Understanding them, it turns out, depends on two very
important physical properties of substances.
Exact Differentials
We have seen that the total differential of U (V , T ) can be expressed as Equation . In general, if a differential can be
4.2.2
expressed as
df (x, y) = P dx + Q dy
In order to illustrate this concept, consider p(V , T ) using the ideal gas law.
RT
p =
V
Solution
Let’s confirm!
4.2.1 https://chem.libretexts.org/@go/page/84309
1 RT ? 1 R
[ (− )] = [ ( )]
2
∂T V ∂V V
V T
R ✓ R
(− ) = (− )
2 2
V V
The differentials of all of the thermodynamic functions that are state functions will be exact. Heat and work are not exact
differential and dw and dq are called inexact differentials instead.
This page titled 4.2: Total and Exact Differentials is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
4.2.2 https://chem.libretexts.org/@go/page/84309
4.3: Compressibility and Expansivity
Isothermal Compressibility (κ ) T
A very important property of a substance is how compressible it is. Gases are very compressible, so when subjected to high
pressures, their volumes decrease significantly (think Boyle’s Law!) Solids and liquids however are not as compressible. However,
they are not entirely incompressible! High pressure will lead to a decrease in volume, even if it is only slight. And, of course,
different substances are more compressible than others.
To quantify just how compressible substances are, it is necessary to define the property. The isothermal compressibility is defined
by the fractional differential change in volume due to a change in pressure.
1 ∂V
κT ≡ − ( ) (4.3.1)
V ∂p
T
The negative sign is important in order to keep the value of κ positive, since an increase in pressure will lead to a decrease in
T
volume. The 1/V term is needed to make the property intensive so that it can be tabulated in a useful manner.
As was the case with the compressibility factor, the 1/V term is needed to make the property intensive, and thus able to be
tabulated in a useful fashion. In the case of expansion, volume tends to increase with increasing temperature, so the partial
derivative is positive.
Deriving an Expression for a Partial Derivative (Type I): The reciprocal rule
Consider a system that is described by three variables, and for which one can write a mathematical constraint on the variables
F (x, y, z) = 0
Under these circumstances, one can specify the state of the system varying only two parameters independently because the
third parameter will have a fixed value. As such one could define two functions: z(x, y) and y(x, z).
This allows one to write the total differentials for dz and dy as follows
∂z ∂z
dz = ( ) dx + ( ) dy (4.3.3)
∂x ∂y
y x
and
∂y ∂y
dy = ( ) dx + ( ) dz (4.3.4)
∂x z
∂z x
∂z ∂z ∂y ∂z ∂y
=( ) dx + ( ) ( ) dx + ( ) ( ) dz (4.3.6)
∂x ∂y ∂x ∂y ∂z
y x z x x
4.3.1 https://chem.libretexts.org/@go/page/84310
If the system undergoes a change following a pathway where x is held constant (dx = 0 ), this expression simplifies to
∂z ∂y
dz = ( ) ( ) dz
∂y ∂z
x x
This reciprocal rule is very convenient in the manipulation of partial derivatives. But it can also be derived in a straight-
forward, albeit less rigorous, manner. Begin by writing the total differential for z(x, y) (Equation 4.3.3):
∂z ∂z
dz = ( ) dx + ( ) dy
∂x ∂y
y x
Noting that
dz ∣
∣ =1
dz ∣x
dx ∣
∣ =0
dz ∣ x
and
dy ∣ ∂y
∣ =( )
∣
dz x ∂z
x
or
∂z 1
( ) =
∂y ∂y
z
( )
∂z
x
This “formal” method of partial derivative manipulation is convenient and useful, although it is not mathematically rigorous.
However, it does work for the kind of partial derivatives encountered in thermodynamics because the variables are state
variables and the differentials are exact.
Deriving an Expression for a Partial Derivative (Type II): The Cyclic Permutation Rule
This alternative derivation follow the initial steps in the derivation above to Equation 4.3.6:
∂z ∂z ∂y ∂z ∂y
dz = ( ) dx + ( ) ( ) dx + ( ) ( ) dz
∂x ∂y ∂x ∂y ∂z
y x z x x
If the system undergoes a change following a pathway where z is held constant (dz = 0 ), this expression simplifies to
∂z ∂z ∂y
0 =( ) dy + ( ) ( ) dx
∂x ∂y ∂x
y x z
4.3.2 https://chem.libretexts.org/@go/page/84310
And so for and changes in which dx ≠ 0
∂z ∂z ∂y
( ) = −( ) ( )
∂x ∂y ∂x
y x z
This cyclic permutation rule is very convenient in the manipulation of partial derivatives. But it can also be derived in a
straight-forward, albeit less rigorous, manner. As with the derivation above, we wegin by writing the total differential of
z(x, y)
∂z ∂z
dz = ( ) dx + ( ) dy
∂x ∂y
y x
Note that
dz ∣
∣ =0
dx ∣z
dx ∣
∣ =1
dx ∣ z
and
dy ∣ ∂y
∣ =( )
∣
dx z ∂x
z
This type of transformation is very convenient, and will be used often in the manipulation of partial derivatives in
thermodynamics.
in terms of derivatives of thermodynamic functions using the definitions in Equations 4.3.1 and 4.3.2.
Solution
Substituting Equations 4.3.1 and 4.3.2 into the Equation 4.3.9
1 ∂V
( )
α V ∂T
p
=
κT 1 ∂V
− ( )
V ∂p
T
4.3.3 https://chem.libretexts.org/@go/page/84310
Simplifying (canceling the 1/V terms and using transformation Type I to invert the partial derivative in the denominator)
yields
α ∂V ∂p
= −( ) ( )
κT ∂T ∂V
p T
This page titled 4.3: Compressibility and Expansivity is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
4.3.4 https://chem.libretexts.org/@go/page/84310
4.4: The Joule Experiment
Going back to the expression for changes in internal energy that stems from assuming that U is a function of V and T (or U (V , T )
for short)
∂U ∂U
dU = ( ) dV + ( ) dT
∂V ∂T
T V
one quickly recognizes one of the terms as the constant volume heat capacity, C . And so the expression can be re-written
V
∂U
dU = ( ) dV + CV dT
∂V
T
But what about the first term? The partial derivative is a coefficient called the “internal pressure”, and given the symbol π .
T
∂U
πT = ( )
∂V
T
James Prescott Joule (1818-1889) recognized that πT should have units of pressure (Energy/volume = pressure) and designed an
experiment to measure it.
He immersed two copper spheres, A and B, connected by a stopcock. Sphere A is filled with a sample of gas while sphere B was
evacuated. The idea was that when the stopcock was opened, the gas in sphere A would expand (ΔV > 0 ) against the vacuum in
sphere B (doing no work since p = 0 . The change in the internal energy could be expressed
ext
dU = πT dV + CV dT
dU = dq + dw
πT dV + CV dT = dq + dw
and since dw = 0
πT dV + CV dT = dq
Joule concluded that dq = 0 (and dT = 0 as well) since he did not observe a temperature change in the water bath which could
only have been caused by the metal spheres either absorbing or emitting heat. And because dV > 0 for the gas that underwent the
expansion into an open space, π must also be zero! In truth, the gas did undergo a temperature change, but it was too small to be
T
detected within his experimental precision. Later, we (once we develop the Maxwell Relations) will show that
∂U ∂p
( ) = T( ) −p (4.4.1)
∂V ∂T
T V
4.4.1 https://chem.libretexts.org/@go/page/84311
Application to an Ideal Gas
For an ideal gas p = RT /V , so it is easy to show that
∂p R
( ) = (4.4.2)
∂T V
V
So while Joule’s observation was consistent with limiting ideal behavior, his result was really an artifact of his experimental
uncertainty masking what actually happened.
so
∂p R
( ) = (4.4.5)
∂T V −b
V
and
∂U R
( ) =T −p (4.4.6)
∂V V −b
T
Substitution of the expression for p (Equation 4.4.4) into this Equation 4.4.6
∂U a
( ) =
2
∂V V
T
And so the internal pressure can be expressed entirely in terms of measurable properties
∂U α
( ) =T −p
∂V κT
T
This page titled 4.4: The Joule Experiment is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
4.4.2 https://chem.libretexts.org/@go/page/84311
4.5: The Joule-Thomson Effect
In 1852, working with William Thomson (who would later become Lord Kelvin), Joule conducted an experiment in which they
pumped gas at a steady rate through a lead pipe that was cinched to create a construction. On the upstream side of the constriction,
the gas was at a higher pressure than on the downstream side of the constriction. Also, the temperature of the gas was carefully
monitored on either side of the construction. The cooling that they observed as the gas expanded from a high pressure region to a
lower pressure region was extremely important and lead to a common design of modern refrigerators.
Not all gases undergo a cooling effect upon expansion. Some gases, such as hydrogen and helium, will experience a warming effect
upon expansion under conditions near room temperature and pressure. The direction of temperature change can be determined by
measuring the Joule-Thomson coefficient, μ . This coefficient has the definition
JT
∂T
μJT ≡ ( )
∂p
H
Schematically, the Joule-Thomson coefficient can be measured by measuring the temperature drop or increase a gas undergoes for
a given pressure drop (Figure 4.5.1). The apparatus is insulated so that no heat can be transferred in or out, making the expansion
isenthalpic.
Figure 4.5.1 :
The typical behavior of the Joule-Thomson coefficient can be summarized in Figure 4.5.2. At the combinations of T and p for
which μ > 0 (inside the shaded region), the sample will cool upon expansion. At those p and T conditions outside of the shaded
JT
region, where μ < 0 , the gas will undergo a temperature increase upon expansion. And along the boundary, a gas will undergo
JT
neither a temperature increase not decrease upon expansion. For a given pressure, there are typically two temperatures at which
μJT changes sign. These are the upper and lower inversion temperatures.
Figure 4.5.2 : The typical behavior of the Joule-Thomson coefficient at different temperatures and pressures.
Using the tools of mathematics, it is possible to express the Joule-Thomson coefficient in terms of measurable properties. Consider
enthalpy as a function of pressure and temperature: H (p, T ). This suggests that the total differential dH can be expressed
∂H ∂H
dH = ( ) dp + ( ) dT (4.5.1)
∂p ∂T
T p
It will be shown later (again, once we develop the Maxwell Relations) that
4.5.1 https://chem.libretexts.org/@go/page/84312
∂H ∂V
( ) dp = −T ( ) +V
∂p ∂T
T p
So
dH = V (1 − T α)dP + Cp dT
1
dH = V (1 − T ) dP + Cp dT
T
which causes the first term to vanish. So for constant enthalpy expansion (dH = 0 ), there can be no change in temperature (
dT = 0 ). This will mean that gases will only show non-zero values for μ only because they deviate from ideal behavior!
JT
Example 4.5.1:
Solution
Using the total differential for H (p, T ) (Equation 4.5.1):
∂H ∂H
dH = ( ) dp + ( ) dT
∂p ∂T
T p
Noting that
dH ∣
∣ =0
dp ∣H
dp ∣
∣ =1
dp ∣ H
and
dT ∣ ∂T
∣ =( )
dp ∣ H
∂p
H
so
∂H ∂H ∂T
0 =( ) +( ) ( )
∂p ∂T ∂p
T p H
∂H
( ) = Cp
∂T
p
4.5.2 https://chem.libretexts.org/@go/page/84312
∂T
( ) = μJT
∂p
H
To get
0 = V (1 − T α) + Cp μJT
This page titled 4.5: The Joule-Thomson Effect is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
4.5.3 https://chem.libretexts.org/@go/page/84312
4.6: Useful Definitions and Relationships
In this chapter (and in the previous chapter), several useful definitions have been stated.
and
∂H
Cp ≡ ( )
∂T
p
or
∂V
( ) =V α
∂T
p
Isothermal Compressibility:
1 ∂V
κT ≡ − ( )
V ∂p
T
or
∂V
( ) = −V κT
∂p
T
and
∂H ∂V
( ) = −T ( ) −p
∂p ∂T
T p
Together, these relationships and definitions make a powerful set of tools that can be used to derive a number of very useful
expressions.
Solution 1:
4.6.1 https://chem.libretexts.org/@go/page/84313
Begin by using the total differential of H (p, T ):
∂H ∂H
dH = ( ) dp + ( ) dT
∂p ∂T
T p
Divide by dV and constrain to constant T (to generate the partial of interest on the left):
0
dH ∣ ∂H dp ∣ ∂H dT ∣
∣ =( ) ∣ + ( ) ∣
dV ∣ T
∂p dV ∣ T
∂T dV ∣ T
T p
The last term on the right will vanish (since dT =0 for constant T ). After converting to partial derivatives
∂H ∂H ∂p
( ) =( ) ( ) (4.6.1)
∂V ∂p ∂V
T T T
This result is simply a demonstration of the “chain rule” on partial derivatives! But now we are getting somewhere. We can
∂H
now substitute for ( ) using our “toolbox of useful relationships”:
∂V T
∂H ∂V ∂p
( ) = [−T ( ) +V ] ( )
∂V ∂T ∂V
T p T
Using the cyclic permutation rule (Transformation Type II), the middle term of Equation 4.6.2 can be simplified
And now all of the partial derivatives on the right can be expressed in terms of α and κT (along with T and V , which are also
“measurable properties”.
∂H α 1
( ) =T +V
∂V κT −V κT
T
or
∂H 1
( ) = (T α − 1)
∂V κT
T
Calculate ΔH for the isothermal compression of ethanol which will decrease the molar volume by 0.010 L/mol at 300 K.
(For ethanol, α = 1.1 × 10 K and κ = 7.9 × 10 atm ).
−3 −1
T
−5 −1
Solution
Integrating the total differential of H at constant temperature results in
∂H
ΔH = ( ) ΔV
∂V
T
4.6.2 https://chem.libretexts.org/@go/page/84313
∂H 1
( ) = (T α − 1)
∂V κT
T
so
1 −3 −1
ΔH = [ ((300 K)(1.1 × 10 K ) − 1)] (−0.010 L/mol)
−5 −1
7.9 × 10 atm
atm L 8.314 J
ΔH = (84.81 ) ( ) = 9590 J/mol
mol
0.8206 atm L
conversion factor
Contributors
Patrick E. Fleming (Department of Chemistry and Biochemistry; California State University, East Bay)
This page titled 4.6: Useful Definitions and Relationships is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated
by Patrick Fleming.
4.6.3 https://chem.libretexts.org/@go/page/84313
4.E: Putting the First Law to Work (Exercises)
Q4.1
Given the relationship
∂U ∂p
(− ) = T (− ) −p (4.E.1)
∂V ∂T
T V
show that
∂U
(− ) =0 (4.E.2)
∂V
T
Q4.2
Determine if the following differential is exact, and if so, find the function z(x, y) that satisfies the expression.
2
dz = 4xy dz + 2 x dy (4.E.3)
Q4.3
For a van der Waals gas,
2
∂U an
( ) =( ) (4.E.4)
2
∂V V
T
if C = 3/2R . Use the expression to calculate the temperature change for 1.00 mol of Xe (a = 4.19 atm L2 mol -2) expanding
V
Q4.4
Given the following data, calculate the change in volume for 50.0 cm3 of
a. neon and
b. copper
due to an increase in pressure from 1.00 atm to 0.750 atm at 298 K.
Ne 1.00 atm-1
Q4.5
Consider a gas that follows the equation of state
nRT
p = (4.E.6)
V − nb
4.E.1 https://chem.libretexts.org/@go/page/84519
V
μJT = (T α − 1) (4.E.7)
Cp
Q4.6
Given
∂H ∂V
( ) = −T ( ) +V (4.E.8)
∂p ∂T
T p
in terms of measurable properties. Use your result to calculate the change in the internal energy of 18.0 g of water when the
pressure is increased from 1.00 atm to 20.0 atm at 298 K.
Q4.7
Derive an expression for
∂U
( ) (4.E.10)
∂T
p
Finish by dividing by dT and constraining to constant pressure. Make substitutions for the measurable quantities, and solve for
∂U
( ) . (4.E.12)
∂T
p
Q4.8
Derive an expression for the difference between C and C in terms of the internal pressure, α , p and V . Using the definition for
p V
∂H ∂U ∂V
( ) =( ) + p( ) (4.E.13)
∂T ∂T ∂T
p p p
Now, find an expression for by starting with U (V , T ) and writing an expression for the total differential dU (V , T ) .
Divide this expression by dp and constrain to constant T . Substitute this into the previous expressions and solve for
∂G ∂U
( ) −( ) . (4.E.14)
∂T ∂T
p V
Q4.9
Evaluate the expression you derived in problem 8 for an ideal, assuming that the internal pressure of an ideal gas is zero.
This page titled 4.E: Putting the First Law to Work (Exercises) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by Patrick Fleming.
4.E.2 https://chem.libretexts.org/@go/page/84519
4.S: Putting the First Law to Work (Summary)
Learning Objectives
After mastering the material covered in this chapter, one will be able to:
1. Express the total differential of a thermodynamic function in terms of partial differentials involving two independent state
variables:
∂U ∂U
dU = ( ) dV + ( ) dT (4.S.1)
∂V ∂T
T V
and
∂z 1
( ) =
∂y ∂y
z
( )
∂z
x
4. Define and describe the meaning of the isobaric thermal expansivity coefficient (α ) and the isothermal compressibility
coefficient (κ ).
T
5. Derive expressions for α and κ for gases based on an assumed equation of state.
T
6. Define internal pressure and describe the experiment Joule used to attempt to measure it.
7. Calculate a value for the internal pressure based on α and κ for a given substance.
T
8. Derive an expression for the internal pressure of a gas based on an assumed equation of state, given
∂U α
( ) =T −p
∂V κT
T
∂H 1
( ) = (T α − 1)
∂V κT
T
12. Demonstrate that the Joule-Thomson coefficient (μ for an ideal gas is zero.
JT
13. Derive expressions for the temperature and pressure dependence of enthalpy and internal energy in terms of measurable
properties. Use these expressions to calculate changes in enthalpy and internal energy for specific substances based on the
values of those measurable properties when the temperature or pressure is changed.
4.S.1 https://chem.libretexts.org/@go/page/84314
References
1. Encyclopedia Brittannica. (2016). Retrieved March 15, 2016, from James Prescott Joule: English Physicist:
http://www.britannica.com/biography/...Prescott-Joul
This page titled 4.S: Putting the First Law to Work (Summary) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by Patrick Fleming.
4.S.2 https://chem.libretexts.org/@go/page/84314
CHAPTER OVERVIEW
This page titled 5: The Second Law is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick Fleming.
1
5.1: Introduction to the Second Law
Rudolph Clausius is kind enough in his 1879 work “The Mechanical Theory of Heat” (Clausius, 1879) to indicate where we have
been in our discussion of thermodynamics, as well as where we are going.
“The fundamental laws of the universe which correspond to the two fundamental theorems of the mechanical
theory of heat:
1. The energy of the universe is constant.
2. The entropy of the universe tends to a maximum.”
― Rudolf Clausius, The Mechanical Theory Of Heat
The second law of thermodynamics, which introduces us to the topic of entropy, is amazing in how it constrains what we can
experience and what we can do in the universe. As Sean M. Carroll, a CalTech Theoretical physicist, suggests in a 2010 interview
with Wired Magazine (Biba, 2010),
I’m trying to understand how time works. And that’s a huge question that has lots of
different aspects to it. A lot of them go back to Einstein and spacetime and how we
measure time using clocks. But the particular aspect of time that I’m interested in is the
arrow of time: the fact that the past is different from the future. We remember the past but
we don’t remember the future. There are irreversible processes. There are things that
happen, like you turn an egg into an omelet, but you can’t turn an omelet into an egg.
We, as observers of nature, are time travelers. And the constraints on what we can observe as we move through time step from the
second law of thermodynamics. But more than just understanding what the second law says, we are interested in what sorts of
processes are possible. And even more to the point, what sorts of processes are spontaneous.
A spontaneous process is one that will occur without external forces pushing it. A process can be spontaneous even if it happens
very slowly. Unfortunately, Thermodynamics is silent on the topic of how fast processes will occur, but is provides us with a
powerful toolbox for predicting which processes will be spontaneous. But in order to make these predictions, a new thermodynamic
law and variable is needed since the first law (which defined ΔU and ΔH ) is insufficient.
Consider the following processes:
+ −
N aOH (s) → N a (aq) + OH (aq)
with ΔH < 0
−
N aH C O3 (s) → N a + (aq) + H C O (aq)
3
with ΔH > 0
Both reactions will occur spontaneously, but one is exothermic and the other endothermic. So while it is intuitive to think that an
exothermic process will be spontaneous, there is clearly more to the picture than simply the release of energy as heat when it comes
to making a process spontaneous. The Carnot cycle because a useful thought experiment to explore to help to answer the question
of why a process is spontaneous.
Contributors
Patrick E. Fleming (Department of Chemistry and Biochemistry; California State University, East Bay)
This page titled 5.1: Introduction to the Second Law is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
5.1.1 https://chem.libretexts.org/@go/page/84316
5.2: Heat Engines and the Carnot Cycle
Heat Engines
Sadi Carnot (1796 – 1832) (Mendoza, 2016), a French physicist and engineer was very interested in the improvement of steam
engines to perform the tasks needed by modern society.
Unfortunately, such a device is impractical. As it turns out, nature prevents the complete conversion of energy into work with
perfect efficiency. This leads to an important statement of the Second Law of Thermodynamics.
It is impossible to convert heat into an equivalent amount of work without some other
changes occurring in the universe.
As such, a more reasonable picture of the heat engine is one which will allow for losses of energy to the surroundings.
The fraction of energy supplied to the engine that can be converted to work defines the efficiency of the engine.
and all of the energy lost to the surroundings also occurs isothermally and reversibly at temperature T . In order to insure this, the
l
5.2.1 https://chem.libretexts.org/@go/page/84317
Thus, the cycle consists of four reversible legs, two of which are isothermal, and two of which are adiabatic.
I. Isothermal expansion from p1 and V1 to p2 and V2 at Th.
II. Adiabatic expansion from p2, V2, Th to p3, V3, Tl.
III. Isothermal compression from p3 and V3 to p4 and V4 at Tl.
IV. Adiabatic compression from p4, V4, Tl to p1, V1, Th.
Plotted on a pressure-volume diagram, the Carnot cycle looks as follows:
Because this is a closed cycle (the ending state is identical initial state) any state function must have a net change of zero as the
system moves around the cycle. Furthermore, the efficiency of the engine can be expressed by the net amount of work the engine
produces per unit of heat supplied to power the engine.
wnet
ϵ=
qh
In order to examine this expression, it is useful to write down expressions fo the heat and work flow in each of the four legs of the
engine cycle.
II 0 CV(Tl – Th)
IV 0 CV(Th – Tl)
5.2.2 https://chem.libretexts.org/@go/page/84317
The total amount of work done is given by the sum of terms in the thirst column. Clearly the terms for the two adiabatic legs cancel
(as they have the same magnitude, but opposite signs.) So the total work done is given by
V2 V4
wtot = nRTh ln( ) + nRTl ln( )
V1 V3
The efficiency of the engine can be defined as the total work produced per unit of energy provided by the high temperature
reservoir.
wtot
ϵ=
qh
or
V2 V4
nRTh ln( ) + nRTl ln( )
V1 V3
ϵ= (5.2.1)
V2
nRTh ln( )
V1
That expression has a lot of variables, but it turns out that it can be simplified dramatically. It turns out that by the choice of
pathways connecting the states places a very important restriction on the relative values of V1, V2, V3 and V4. To understand this,
we must consider how the work of adiabatic expansion is related to the initial and final temperatures and volumes. In Chapter 3, it
was shown that the initial and final temperatures and volumes of an adiabatic expansion are related by
CV /R CV /R
Vi T = Vf T
i f
or
CV /R
Vi Tf
=( )
Vf Ti
Using the adiabatic expansion and compression legs (II and IV), this requires that
CV /R
V2 Th
=( )
V2 Tl
and
CV /R
V4 Tl
=( )
V1 Th
Since the second terms are reciprocals of one another, the first terms must be as well!
V2 V1
=
V2 V4
This is very convenient! It is what allows for the simplification of the efficiency expression (Equation 5.2.1) becomes
V2 V2
nR Th ln( ) + nR Tl ln( )
V1 V1
ϵ=
V2
nR Th ln( )
V1
5.2.3 https://chem.libretexts.org/@go/page/84317
Tg − Tl
ϵ= (5.2.2)
Th
This expression gives the maximum efficiency and depends only on the high and low temperatures!
Also, it should be noted that the heat engine can be run backwards. By providing work to the engine, it can be forces to draw heat
from the low temperature reservoir and dissipate it into the high temperature reservoir. This is how a refrigerator or heat pump
works. The limiting efficiency of such a device can also be calculated using the temperatures of the hot can cold reservoirs.
Example 5.2.1:
What is the maximum efficiency of a freezer set to keep ice cream at a cool -10 oC, which it is operating in a room that is
25oC? What is the minimum amount of energy needed to remove 1.0 J from the freezer and dissipate it into the room?
Solution
The efficiency is given by Equation 5.2.2 and converting the temperatures to an absolute scale, the efficiency can be calculated
as
298 K − 263 K
ϵ=
298 K
So
1.0 J = 0.1174(w)
or
w = 8.5 J
It is interesting to note that any arbitrary closed cyclical process can be described as a sum of infinitesimally small Carnot cycles,
and so all of the conclusions reached for the Carnot cycle apply to any cyclical process.
Contributors
Patrick E. Fleming (Department of Chemistry and Biochemistry; California State University, East Bay)
This page titled 5.2: Heat Engines and the Carnot Cycle is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated
by Patrick Fleming.
5.2.4 https://chem.libretexts.org/@go/page/84317
5.3: Entropy
In addition to learning that the efficiency of a Carnot engine depends only on the high and low temperatures, more interesting
things can be derived through the exploration of this system. For example, consider the total heat transferred in the cycle:
V2 V4
qtot = nRTh ln( ) − nRTl ln( )
V1 V3
It is clear that the two terms do not have the same magnitude, unless T = T . This is sufficient to show that q is not a state
h l
function, since it’s net change around a closed cycle is not zero (as any value of a state function must be.) However, consider what
happens when the sum of q/T is considered:
V4 V4
nR Th ln( ) nR Tl ln( )
q V3 V3
∑ = −
T Th Tl
V4 V4
= nR ln( ) − nR ln( )
V3 V3
=0
This is the behavior expected for a state function! It leads to the definition of entropy in differential form,
dqrev
dS ≡
T
In general, dq will be larger than dq (since the reversible pathway defines the maximum heat flow.) So, it is easy to calculate
rev
entropy changes, as one needs only to define a reversible pathway that connects the initial and final states, and then integrate dq/T
over that pathway. And since \(\Delta S\) is defined using q for a reversible pathway, ΔS is independent of the actual path a system
follows to undergo a change.
This page titled 5.3: Entropy is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick Fleming.
5.3.1 https://chem.libretexts.org/@go/page/84318
5.4: Calculating Entropy Changes
Entropy changes are fairly easy to calculate so long as one knows initial and final state. For example, if the initial and final volume
dq
are the same, the entropy can be calculated by assuming a reversible, isochoric pathway and determining an expression for . That T
term can then be integrated from the initial condition to the final conditions to determine the entropy change.
Isothermal Changes
If the initial and final temperatures are the same, the most convenient reversible path to use to calculate the entropy is an isothermal
pathway. As an example, consider the isothermal expansion of an ideal gas from V to V . As was derived in Chapter 3,
1 2
dV
dq = nRT
V
So dq/T is given by
dq dV
= nR
T V
and so
V2
dq dV V2
ΔS = ∫ = nR ∫ = nR ln( ) (5.4.1)
T V1
V V1
Solution
Recognizing that this is an isothermal process, we can use Equation 5.4.1
V2
ΔS = nR ln( )
V1
44.8 L
= (1.00 mol)(8.314J/(mol K)) ln( )
22.4 L
= 5.76 J/K
Isobaric Changes
For changes in which the initial and final pressures are the same, the most convenient pathway to use to calculate the entropy
change is an isobaric pathway. In this case, it is useful to remember that
dq = nCp dT
So
dq dT
= nCp
T T
Integration from the initial to final temperature is used to calculate the change in entropy. If the heat capacity is constant over the
temperature range
T2 T2
dq dT T2
∫ = nCp ∫ = nCp ln( )
T1
T T1
T T1
If the temperature dependence of the heat capacity is known, it can be incorporated into the integral. For example, if Cp can be
expressed as
5.4.1 https://chem.libretexts.org/@go/page/84319
c
Cp = a + bT +
2
T
∫ = n∫ dT
T1
T T1
T
which simplifies to
T2
a c
ΔS = n ∫ ( + bT + ) dT
3
T1 T T
or
T2 c 1 1
ΔS = n [a ln( ) + b(T2 − T1 ) − ( − )]
2 2
T1 2 T T
2 1
Isochoric Changes
Similarly to the cast of constant pressure, it is fairly simple to calculate ΔS . Since
dq = nCV dt
dq
is given by
T
dq dT
= nCV
T T
T2
ΔS = nCv ln( )
T1
Adiabatic Changes
The easiest pathway for which to calculate entropy changes is an adiabatic pathway. Since dq = 0 for an adiabatic change, then
dS = 0 as well.
Phase Changes
The entropy change for a phase change at constant pressure is given by
q ΔHphase
ΔS = = (5.4.2)
T T
Solution
This is a phase transition at constant pressure (assumed) requiring Equation 5.4.2:
(1 mol)(6010 J/mol)
ΔS =
273 K
= 22 J/K
5.4.2 https://chem.libretexts.org/@go/page/84319
This page titled 5.4: Calculating Entropy Changes is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
5.4.3 https://chem.libretexts.org/@go/page/84319
5.5: Comparing the System and the Surroundings
It is oftentimes important (for reasons that will be discussed in the next section) to calculate both the entropy change of the system
as well as that of the surroundings. Depending on the size of the surroundings, they can provide or absorb as much heat as is
needed for a process without changing temperature. As such, it is oftentimes a very good approximation to consider the changes to
the surroundings as happening isothermally, even though it may not be the case for the system (which is generally smaller.)
Example 5.5.1
Consider 18.02 g (1.00 mol) of ice melting at 273 K in a room that is 298 K. Calculate DS for the ice, the surrounding room,
and of the universe. (DHfus = 6.01 kJ/mol)
Solution
For the process under constant pressure: q ice = −qroom :
Example 5.5.2
A 10.0 g piece of metal (C = 0.250 J/g °C) initially at 95 °C is placed in 25.0 g of water initially at 15 °C in an insulated
container. Calculate the final temperature of the metal and water once the system has reached thermal equilibrium. Also,
calculate the entropy change for the metal, the water, and the entire system.
Solution
Heat will be transferred from the hot metal to the cold water. Since it has nowhere else to go, the final temperature can be
calculated from the expression
qw = −qm
where q is the heat absorbed by the water, and q is the heat lost by the metal. And since
w m
q = mC ΔT
it follows that
Tf = 16.9 °C .
5.5.1 https://chem.libretexts.org/@go/page/84320
To get the entropy changes, use the expression:
Tf
ΔS = m Cp ln( )
Ti
= 0.689 J/K
= −0.596 J/K
Note: The total entropy change is positive, suggesting that this will be a spontaneous process. This should make some sense
since one expects heat to flow from the hot metal to the cool water rather than the other way around. Also, note that the sign of
the entropy change is positive for the part of the system that is absorbing the heat, and negative for the part losing the heat.
Adiabatic 0
drev V2
Isothermal nR ln( ) *
T V1
Tf
Isobaric mCp ln( ) qsys
Ti
ΔSsurr =
Tsurr
Tf
Isochoric mCV ln( )
Ti
ΔHphase
Phase Change
T
And
This calculation is important as ΔS provides the criterion for spontaneity for which we were searching from the outset. This
univ
5.5.2 https://chem.libretexts.org/@go/page/84320
Clausius Inequality
The Second Law can be summed up in a very simple mathematical expression called the Clausius Inequality.
ΔSuniverse ≥ 0
which must be true for any spontaneous process. It is not the most convenient criterion for spontaneity, but it will do for now.
In the next chapter, we will derive a criterion which is more useful to us as chemists, who would rather focus on the system
itself rather than both the system and its surroundings. Another statement of the Clausius theorem is
dq
∮ ≥0
T
with the only condition of the left hand side equaling zero is if the system transfers all heat reversibly.
Contributors
Patrick E. Fleming (Department of Chemistry and Biochemistry; California State University, East Bay)
This page titled 5.5: Comparing the System and the Surroundings is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by Patrick Fleming.
5.5.3 https://chem.libretexts.org/@go/page/84320
5.6: Entropy and Disorder
A common interpretation of entropy is that it is somehow a measure of chaos or randomness. There is some utility in that concept.
Given that entropy is a measure of the dispersal of energy in a system, the more chaotic a system is, the greater the dispersal of
energy will be, and thus the greater the entropy will be. Ludwig Boltzmann (1844 – 1906) (O'Connor & Robertson, 1998)
understood this concept well, and used it to derive a statistical approach to calculating entropy. Boltzmann proposed a method for
calculating the entropy of a system based on the number of energetically equivalent ways a system can be constructed.
Boltzmann proposed an expression, which in its modern form is:
S = kb ln(W ) (5.6.1)
This rather famous equation is etched on Boltzmann’s grave marker in commemoration of his profound contributions to the science
of thermodynamics (Figure 5.6.1).
Example 5.6.1:
Calculate the entropy of a carbon monoxide crystal, containing 1.00 mol of CO, and assuming that the molecules are randomly
oriented in one of two equivalent orientations.
Solution
Using the Boltzmann formula (Equation 5.6.1):
S = nK ln(W )
= 5.76 J/K
This page titled 5.6: Entropy and Disorder is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
5.6.1 https://chem.libretexts.org/@go/page/84321
5.7: The Third Law of Thermodynamics
One important consequence of Botlzmann’s proposal is that a perfectly ordered crystal (i.e. one that has only one energetic
arrangement in its lowest energy state) will have an entropy of 0. This makes entropy qualitatively different than other
thermodynamic functions. For example, in the case of enthalpy, it is impossible have a zero to the scale without setting an arbitrary
reference (which is that the enthalpy of formation of elements in their standard states is zero.) But entropy has a natural zero! It is
the state at which a system has perfect order. This also has another important consequence, in that it suggests that there must also
be a zero to the temperature scale. These consequences are summed up in the Third Law of Thermodynamics.
where C is the heat capacity. An entropy value determined in this manner is called a Third Law Entropy.
Naturally, the heat capacity will have some temperature dependence. It will also change abruptly if the substance undergoes a phase
change. Unfortunately, it is exceedingly difficult to measure heat capacities very near zero K. Fortunately, many substances follow
the Debye Extrapolation in that at very low temperatures, their heat capacities are proportional to T3. Using this assumption, we
have a temperature dependence model that allows us to extrapolate absolute zero based on the heat capacity measured at as low a
temperature as can be found.
Example 5.7.1
SiO2 is found to have a molar heat capacity of 0.777 J mol-1 K-1 at 15 K (Yamashita, et al., 2001). Calculate the molar entropy
of SiO2 at 15 K.
Solution
Using the Debye model, the heat capacity is given by
This page titled 5.7: The Third Law of Thermodynamics is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated
by Patrick Fleming.
5.7.1 https://chem.libretexts.org/@go/page/84496
5.8: Adiabatic Compressibility
In Chapter 4, we learned about the isothermal compressibility, κ , which is defined as
T
1 ∂V
κT = − ( )
V ∂p
T
κT is a very useful quantity, as it can be measured for many different substances and tabulated. Also, as we will see in the next
chapter, it can be used to evaluate several different partial derivatives involving thermodynamic variables.
In his seminal work, Philosophiae Naturalis Principia Mathematica (Newton, 1723), Isaac Newton (1643 - 1727) (Doc) calculated
the speed of sound through air, assuming that sound was carried by isothermal compression waves. His calculated value of 949 m/s
was about 15% smaller than experimental determinations. He accounted for the difference by pointing to “non-ideal effects”. But it
turns out that his error, albeit an understandable one (since sound waves do not appear to change bulk air temperatures) was that the
compression waves are adiabatic, rather than isothermal. As such, there are small temperature oscillations that occur due to the
adiabatic compression followed by expansion of the gas carrying the sound waves. The oversight was correct by Pierre-Simon
Laplace (1749 – 1827) (O'Connor & Robertson, Pierre-Simon Laplace, 1999).
LaPlace modeled the compression waves using the adiabatic compressibility, κ defined by S
1 ∂V
κS = − ( )
V ∂p
S
it follows that any adiabatic pathway (dq = 0 ) is also isentropic (dS = 0 ), or proceeds at constant entropy.
Derivative Transformation Type II. Applying this, the adiabatic compressibility can be expressed
1 ∂V ∂S
κS = ( ) ( )
V ∂S ∂p
p V
Using a simple chain rule, the partial derivatives can be expanded to get something a little easier to evaluate:
∂S ∂T
( ) ( )
1 ∂T ∂p
V V
κS = (5.8.1)
V ∂S ∂T
( ) ( )
∂T ∂V
p p
∂S Cp
( ) = (5.8.3)
∂T T
p
5.8.1 https://chem.libretexts.org/@go/page/84752
This means that Equation 5.8.1 simplifies to
⎛ ∂T ⎞
( )
CV ⎜ 1 ∂p ⎟
V
κS = ⎜ ⎟
⎜ ⎟
Cp ⎜ V ∂T ⎟
( )
⎝ ∂V ⎠
p
CV 1 ∂T ∂V
κS = ( ( ) ( ) )
Cp V ∂p ∂T
V p
CV 1 ∂V
κS = (− ( ) )
Cp V ∂p
T
CV
κS = κT
Cp
As will be shown in the next chapter, C is always bigger than C , so κ is always smaller than κ .
p V S T
But there is more! We can use this methodology to revisit how pressure affects volume along an adiabat. In order to do this, we
would like to evaluate the partial derivative
∂V
( )
∂p
S
And as before, noting that the relationships in Equations 5.8.2 and 5.8.3, Equation 5.8.4 can be simplified to
∂V CV ∂V ∂T
( ) =− ( ) ( )
∂p Cp ∂T ∂p
S p V
CV ∂V
= ( ) (5.8.5)
Cp ∂p
T
The right-hand derivative is easy to evaluate if we assume a specific equation of state. For an ideal gas,
∂V nRT V
( ) =− =−
∂p p2 p
T
Substitution yields
5.8.2 https://chem.libretexts.org/@go/page/84752
∂V V
γ( ) =−
∂p p
S
which is now looking like a form that can be integrated. Separation of variables yields
dV dP
γ =
V p
or
V2 p2
γ ln( ) = ln( )
V1 p1
or
γ
pV = constant
which is what we previously determined for the behavior of an ideal gas along an adiabat.
Finally, it should be noted that the correct expression for the speed of sound is given by
−−−−
1
vsound = √
ρ κS
where ρ is the density of the medium. For an ideal gas, this expression becomes
−−−−−
γRT
vsound = √
M
where M is the molar mass of the gas. Isaac Newton’s derivation, based on the idea that sound waves involved isothermal
compressions, would produce a result which is missing the factor of γ, accounting for the systematic deviation from experiment
which he observed.
Contributors
Patrick E. Fleming (Department of Chemistry and Biochemistry; California State University, East Bay)
This page titled 5.8: Adiabatic Compressibility is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
5.8.3 https://chem.libretexts.org/@go/page/84752
5.E: The Second Law (Exercises)
Q5.1
What is the minimum amount of work needed to remove 10.0 J of energy from a freezer at -10.0 °C, depositing the energy into a
room that is 22.4 °C?
Q5.2
Consider the isothermal, reversible expansion of 1.00 mol of a monatomic ideal gas (CV = 3/2 R) from 10.0 L to 25.0 L at 298 K.
Calculate q, w, ΔU , ΔH , and ΔS for the expansion.
Q5.3
Consider the isobaric, reversible expansion of 1.00 mol of a monatomic ideal gas (Cp = 5/2 R) from 10.0 L to 25.0 L at 1.00 atm.
Calculate q, w, ΔU , ΔH , and ΔS for the expansion.
Q5.4
Consider the isochoric, reversible temperature increase of 1.00 mol of a monatomic ideal gas (CV = 3/2 R) °Ccupying 25.0 L from
298 K to 345 K. Calculate q, w, ΔU , ΔH , and ΔS for this process.
Q5.5
Consider the adiabatic expansion of 1.00 mol of a monatomic ideal gas (CV = 3/2 R) from 10.0 L at 273 K to a final volume of 45.0
L. Calculate ΔT , q, w, ΔU , ΔH , and ΔS for the expansion.
Q5.6
15.0 g of ice (ΔH = 6.009 kJ/mol) at 0 °C sits in a room that is at 21 °C. The ice melts to form liquid at 0 °C. Calculate the
f us
entropy change for the ice, the room, and the universe. Which has the largest magnitude?
Q5.7
15.0 g of liquid water (Cp = 75.38 J mol-1 °C-1) at 0 °C sits in a room that is at 21 °C. The liquid warms from 0 °C to 21 °C.
Calculate the entropy change for the liquid, the room, and the universe. Which has the largest magnitude?
Q5.8
Calculate the entropy change for taking 12.0 g of H2O from the solid phase (Cp = 36.9 J mol-1 K-1) at -12.0 °C to liquid (Cp = 75.2
J mol-1 K-1) at 13.0 °C. The enthalpy of fusion for water is ΔH = 6.009 kJ/mol.
f us
Q5.9
Using Table T1, calculate the standard reaction entropies (ΔS ) for the following reactions at 298 K.
o
Q5.10
1.00 mole of an ideal gas is taken through a cyclic process involving three steps:
I. Isothermal expansion from V1 to V2 at T1
II. Isochoric heating from, T1 to T2 at V2
III. Adiabatic compression from V2 to V1
a. Graph the process on a V-T diagram.
b. Find q, w, ΔU , and ΔS for each leg. (If you want, you can find ΔH too!)
5.E.1 https://chem.libretexts.org/@go/page/84497
c. Use the fact that ΔS for the entire cycle must be zero (entropy being a state function and all …), determine the relationship
between V1 and V2 in terms of Cv, T1 and T2.
Q5.11
2.00 moles of a monatomic ideal gas (CV = 3/2 R) initially exert a pressure of 1.00 atm at 300.0 K. The gas undergoes the
following three steps, all of which are reversible:
I. isothermal compression to a final pressure of 2.00 atm,
II. Isobaric temperature increase to a final temperature of 400.0 K, and
III. A return to the initial state along a pathway in which
p = a + bT (5.E.1)
where a and b are constants. Sketch the cycle on a pressure-temperature plot, and calculate \(\Delta U\) and \(\Delta S\) for each of
the legs. Are ΔU and ΔS zero for the sum of the three legs?
Q5.12
A 10.0 g piece of iron (C = 0.443 J/g °C) initially at 97.6 °C is placed in 50.0 g of water (C = 4.184 J/g °C) initially at 22.3 °C in an
insulated container. The system is then allowed to come to thermal equilibrium. Assuming no heat flow to or from the
surroundings, calculate
a. the final temperature of the metal and water
b. the change in entropy for the metal
c. the change in entropy for the water
d. the change in entropy for the universe
Q5.13
Considers a crystal of C H F C lBr as having four energetically equivalent orientations for each molecule. What is the expected
residual entropy at 0 K for 2.50 mol of the substance?
Q5.14
A sample of a certain solid is measured to have a constant pressure heat capacity of 0.436 J mol-1 K-1 at 10.0 K. Assuming the
Debeye extrapolation model
3
Cp (T ) = aT (5.E.2)
holds at low temperatures, calculate the molar entropy of the substance at 12.0 K.
This page titled 5.E: The Second Law (Exercises) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
5.E.2 https://chem.libretexts.org/@go/page/84497
5.S: The Second Law (Summary)
Learning Objectives
After mastering the material presented in this chapter, one will be able to:
1. Describe a Carnot engine and derive a relationship for its efficiency of converting heat into work, in terms of the two
temperatures at which the engine operates.
2. Define entropy and be able to calculate entropy changes for systems (and the surroundings) undergoing changes which are
definable as following various pathways, including constant temperature, constant pressure, constant volume, and adiabatic
pathways.
3. Relate entropy to disorder in a crystal based on the number of equivalent orientations a single formula unit may take within the
crystal.
4. State the Third Law of Thermodynamics, and use it to calculate total entropies for substances at a given temperature.
5. Understand how isothermal compressibility differs from adiabatic compressibility and relate that difference to the measurement
of the speed of sound waves traveling through a gas medium.
References
1. Biba, E. (2010, February 26). What is Time? One Physicist Hunts for the Ultimate theory. Wired. Retrieved from
http://www.wired.com/2010/02/what-is-time/
2. Clausius, R. (1879). The Mechanical Theory of Heat. London: McMillan and Company.
3. Doc, T. (n.d.). Isaac Newton. Retrieved April 2, 2016, from FamousScientists.org: http://www.famousscientists.org/isaac-
newton/
4. Mendoza, E. (2016). Sadi Carnot: French engineer and physicist. (Encyclopedia Brittanica, Inc.) Retrieved March 30, 2016,
from Encyclopædia Britannica Online: http://www.britannica.com/biography/...ench-scientist
5. Newton, I. (1723). Philosophiae naturalis principia mathematica. Amstelodamum.
6. O'Connor, J. J., & Robertson, E. F. (1998, September). Retrieved March 30, 2016, from www-groups.dcs.st-
and.ac.uk/~...Boltzmann.html
7. O'Connor, J. J., & Robertson, E. F. (1999). Pierre-Simon Laplace. Retrieved April 2, 2016, from School of Mathematics and
Statistics, University of St Andrews, Scotland: www-groups.dcs.st-and.ac.uk/~...s/Laplace.html
8. Yamashita, I., Tojo, T., Kawaji, H., Atake, T., Linnard, Y., & Richet, P. (2001). Low-temperature heat capacity of sodium
borosilicate glasses at temperatures from 13 K to 300 K. The Journal of Chemical Thermodynamics, 33(5), 535-53.
doi:10.1006/jcht.2000.0744
This page titled 5.S: The Second Law (Summary) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
5.S.1 https://chem.libretexts.org/@go/page/84498
CHAPTER OVERVIEW
This page titled 6: Putting the Second Law to Work is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
1
6.1: Free Energy Functions
In the previous chapter, we saw that for a spontaneous process, ΔS > 0 . While this is a useful criterion for determining
universe
whether or not a process is spontaneous, it is rather cumbersome, as it requires one to calculate not only the entropy change for the
system, but also that of the surroundings. It would be much more convenient if there was a single criterion that would do the job
and focus only on the system. As it turns out, there is.
Since we know that
ΔSuniv ≥ 0
all we need to do is to find an expression for ΔS that can be determined by the changes in the system itself. Fortunately, we
sys
ΔH = qp
Substitution into the above equations yields an expression for the criterion of spontaneity that depends only on variables describing
the changes in the system!
ΔHsys
ΔSuniv ≥ ΔSsys −
T
so
ΔHsys
ΔSsys − ≥0
T
A similar derivation for constant volume processes results in the expression (at constant volume and temperature)
ΔU − T ΔS ≤ 0 (6.1.2)
Equation 6.1.1 is of great use to chemists, as most of chemistry occurs at constant pressure. For geologists, however, who are
interested in processes that occur at very high pressures (say, under the weight of an entire mountain) and expansion is not a
possibility, the constant volume expression of Equation 6.1.1 may be of greater interest.
All of the above arguments can be made for systems in which the temperature is not constant by considering infinitesimal changes.
The resulting expressions are
dH − T dS ≤ 0 (6.1.3)
and
dU − T dS ≤ 0 (6.1.4)
6.1.1 https://chem.libretexts.org/@go/page/84323
The Gibbs and Helmholtz Functions
Equation 6.1.3 suggests a very convenient thermodynamic function to help keep track of both the effects of entropy and enthalpy
changes. This function, the Gibbs function (or Gibbs Free Energy) is defined by
G ≡ H −TS
ΔG = ΔH − Δ(T S)
Or at constant temperature
ΔG = ΔH − T ΔS
And the criterion for a process to be spontaneous is the DG < 0. As such, it should be clear spontaneity is not merely a function the
enthalpy change (although exothermic processes tend to be spontaneous) but also a function of the entropy change, weighted by the
temperature. Going back to an earlier example,
+ −
N aOH (s) → N a (aq) + OH (aq)
with ΔH < 0 .
and
+ −
N aH C O3 (s) → N a (aq) + H C O (aq)
3
with ΔH > 0 .
It is easy to see why both processes are spontaneous. In the first case, the process is exothermic (favorable) and proceeds with an
increase in entropy (also favorable) due to the formation of fragments in the liquid phase (more chaotic) from a very ordered solid
(more ordered). The second reaction is endothermic (unfavorable) but proceeds with an increase in entropy (favorable). So, so long
as the temperature is high enough, the entropy term will overwhelm the enthalpy term and cause the process to be spontaneous. The
conditions for spontaneous processes at constant temperature and pressure can be summarized in Table 6.1.1.
Table 6.1.1: Spontaneity Conditions for a Process under Constant Temperature and Pressure
ΔH ΔS Spontaneous
>0 <0 At no T
A ≡ U −TS
and provides another important criterion for spontaneous processes at constant value and temperature. At constant temperature, the
Helmholtz function can be expressed by
ΔA ≡ ΔU − T ΔS
Based on similar arguments used for the Gibbs function, the Helmholtz function also can be used to predict which processes will be
spontaneous at constant volume and temperature according to Table 6.1.2.
Table 6.1.2: Spontaneity Conditions for a Process under Constant Temperature and Volume
ΔU ΔS Spontaneous?
>0 <0 At no T
6.1.2 https://chem.libretexts.org/@go/page/84323
ΔU ΔS Spontaneous?
their standard states is defined as zero. This allows for two important things to happen. First, ΔG can be measured and tabulated
o
f
elements in their standard states (similarly to how ΔH is defined.) Secondly, tabulated (ΔG ) can be used to calculate standard
o
f
o
f
Example 6.1.1:
Given the following data at 298 K, calculate ΔG at 298 K for the following reaction:
o
Substance ΔG
o
f
(kJ/mol)
C2H4(g) 68.4
C2H6(g) -32.0
Solution
The ΔG values can be used to calculate ΔG for the reaction in exactly the same method as ΔH can be used to calculate a
o
f
o o
f
reaction enthalpy.
o
ΔG = (1 mol)(−32.0 kJ/mol) − (1 mol)(68.4 kJ/mol)
o
ΔG = 100.4 kJ
f
2 (g) is 0 since it is an element in its standard state.
This page titled 6.1: Free Energy Functions is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
6.1.3 https://chem.libretexts.org/@go/page/84323
6.2: Combining the First and Second Laws - Maxwell's Relations
Modeling the dependence of the Gibbs and Helmholtz functions behave with varying temperature, pressure, and volume is
fundamentally useful. But in order to do that, a little bit more development is necessary. To see the power and utility of these
functions, it is useful to combine the First and Second Laws into a single mathematical statement. In order to do that, one notes that
since
dq
dS =
T
dq = T dS
And since
dw = T dS − pdV
for a reversible expansion in which only p-V works is done, it also follows that (since dU = dq + dw ):
dU = T dS − pdV
This is an extraordinarily powerful result. This differential for dU can be used to simplify the differentials for H , A , and G . But
even more useful are the constraints it places on the variables T, S, p, and V due to the mathematics of exact differentials!
Maxwell Relations
The above result suggests that the natural variables of internal energy are S and V (or the function can be considered as U (S, V ) ).
So the total differential (dU ) can be expressed:
∂U ∂U
dU = ( ) dS + ( ) dV
∂S ∂V
V S
and
∂U
( ) = −p (6.2.2)
∂V
S
But the value doesn’t stop there! Since dU is an exact differential, the Euler relation must hold that
∂ ∂U ∂ ∂U
[ ( ) ] =[ ( ) ]
∂V ∂S ∂S ∂V
V S S V
or
∂T ∂p
( ) = −( )
∂V ∂S
S V
This is an example of a Maxwell Relation. These are very powerful relationship that allows one to substitute partial derivatives
when one is more convenient (perhaps it can be expressed entirely in terms of α and/or κ for example.) T
H ≡ U + pV
6.2.1 https://chem.libretexts.org/@go/page/84324
Differentiating (and using the chain rule on d(pV )) yields
dH = dU + pdV + V dp
Making the substitution using the combined first and second laws (dU = T dS– pdV ) for a reversible change involving on
expansion (p-V) work
dH = T dS + V dp (6.2.3)
And much as in the case of internal energy, this suggests that the natural variables of H are S and p. Or
∂H ∂H
dH = ( ) dS + ( ) dV (6.2.4)
∂S p
∂p S
and
∂H
( ) =V (6.2.6)
∂p
S
∂ ∂H ∂ ∂H
[ ( ) ] =[ ( ) ]
∂p ∂S ∂S ∂p
p S p
S
so
∂T ∂V
( ) =( )
∂p ∂S
S p
This is the Maxwell relation on H . Maxwell relations can also be developed based on A and G. The results of those derivations are
summarized in Table 6.2.1..
Table 6.2.1: Maxwell Relations
Function Differential Natural Variables Maxwell Relation
∂T ∂p
U dU = T dS − pdV S, V ( ) = −( )
∂V ∂S
S V
∂T ∂V
H dH = T dS + V dp S, p ( ) = ( )
∂p ∂S
S p
6.2.2 https://chem.libretexts.org/@go/page/84324
Function Differential Natural Variables Maxwell Relation
∂p ∂S
A dA = −pdV − SdT V , T ( ) = ( )
∂T ∂V
V T
∂V ∂S
G dG = V dp − SdT p, T ( ) = −( )
∂T ∂p
p T
The Maxwell relations are extraordinarily useful in deriving the dependence of thermodynamic variables on the state variables of p,
T, and V.
Example 6.2.1
Show that
∂V α
( ) =T −p
∂T κT
p
Solution
Start with the combined first and second laws:
dU = T dS − pdV
Noting that
dU ∣ ∂U
∣ =( )
dV ∣T ∂V
T
T dS ∣ ∂S
∣ =( )
dV ∣ T
∂V
T
dV ∣
∣ =1
dV ∣ T
The result is
∂U ∂S
( ) = T( ) −p
∂V ∂V
T T
to get
∂U ∂p
( ) = T( ) −p
∂V ∂T
T V
and since
∂p α
( ) =
∂T κT
V
It is apparent that
6.2.3 https://chem.libretexts.org/@go/page/84324
∂V α
( ) =T −p
∂T κT
p
Note: How cool is that? This result was given without proof in Chapter 4, but can now be proven analytically using the
Maxwell Relations!
This page titled 6.2: Combining the First and Second Laws - Maxwell's Relations is shared under a CC BY-NC-SA 4.0 license and was authored,
remixed, and/or curated by Patrick Fleming.
6.2.4 https://chem.libretexts.org/@go/page/84324
6.3: ΔA, ΔG, and Maximum Work
The functions A and G are oftentimes referred to as free energy functions. The reason for this is that they are a measure of the
maximum work (in the case of ΔA) or non p-V work (in the case of ΔG) that is available from a process. To show this, consider
the total differentials.
First, consider the differential of A .
dA = dU − T dS − SdT
Substituting the combined first and second laws for dU , but expressing the work term as dw, yields
dA = T dS − dw − T dS − SdT
dA = dw − SdT
dA = dw
Since the only assumption made here was that the change is reversible (allowing for the substitution of T dS for dq), and dw for a
reversible change is the maximum amount of work, it follows that dA gives the maximum work that can be produced from a
process at constant temperature.
Similarly, a simple expression can be derived for dG. Starting from the total differential of G.
Using an expression for dU = dq + dw , where dq = T dS and dw is split into two terms, one (dw pV ) describing the work of
expansion and the other (dw ) describing any other type of work (electrical, stretching, etc.)
e
dU − T dS + dWpV + dWe
dG can be expressed as
dG = dwe
This implies that dG gives the maximum amount of non p-V work that can be extracted from a process.
This concept of dA and dG giving the maximum work (under the specified conditions) is where the term “free energy” comes
from, as it is the energy that is free to do work in the surroundings. If a system is to be optimized to do work in the soundings (for
example a steam engine that may do work by moving a locomotive) the functions A and G will be important to understand. It will,
therefore, be useful to understand how these functions change with changing conditions, such as volume, temperature, and
pressure.
Contributors
Patrick E. Fleming (Department of Chemistry and Biochemistry; California State University, East Bay)
This page titled 6.3: ΔA, ΔG, and Maximum Work is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
6.3.1 https://chem.libretexts.org/@go/page/84325
6.4: Volume Dependence of Helmholtz Energy
If one needs to know how the Helmholtz function changes with changing volume at constant temperature, the following expression
can be used:
V2
∂A
ΔA = ∫ ( ) dV (6.4.1)
V1
∂V
T
But how does one derive an expression for the partial derivative in Equation 6.4.1 ? This is a fairly straight forward process that
begins with the definition of A :
A = U −TS
dA = dU − T dS − SdT (6.4.2)
dU = T dS − pdV (6.4.3)
which assumes:
1. a reversible change and
2. only pV work is being done.
Substituting Equation 6.4.3 into Equation 6.4.2 yields
The natural variables of A are therefore V and T ! So the total differential of A is conveniently expressed as
∂A ∂A
dA = ( ) dV + ( ) dT (6.4.6)
∂V ∂T
T V
∂A
( ) = −S
∂T
V
ΔA = − ∫ p dV
V1
If the pressure is independent of the temperature, it can be pulled out of the integral.
V2
ΔA = −p ∫ dV = −p(V2 − V1 )
V1
ΔA = − ∫ p(V ) dV
V1
Fortunately, this is easy if the substance is an ideal gas (or if some other equation of state can be used, such as the van der Waals
equation.)
6.4.1 https://chem.libretexts.org/@go/page/84326
Example 6.4.1: Ideal Gas Expansion
Calculate ΔA for the isothermal expansion of 1.00 mol of an ideal gas from 10.0 L to 25.0 L at 298 K.
Solution
For an ideal gas,
nRT
p =
V
So
∂A
( ) = −p
∂V
T
becomes
∂A nRT
( ) =−
∂V V
T
becomes
V2
dV
ΔA = −nRT ∫ dT
V1
V
or
V2
ΔA = −nRT ln( )
V1
But further, it is easy to show that the Maxwell relation that arises from the simplified expression for the total differential of A is
∂p ∂S
( ) =( )
∂T ∂V
V T
This particular Maxwell relation is exceedingly useful since one of the terms depends only on p, V , and T . As such it can be
expressed in terms of our old friends, α and κ !T
∂p α
( ) =
∂T κT
V
This page titled 6.4: Volume Dependence of Helmholtz Energy is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by Patrick Fleming.
6.4.2 https://chem.libretexts.org/@go/page/84326
6.5: Pressure Dependence of Gibbs Energy
The pressure and temperature dependence of G is also easy to describe. The best starting place is the definition of G.
G = U + pV − T S (6.5.1)
The differential can be simplified by substituting the combined first and second law statement for dU (consider a reversible process
and pV work only).
This suggests that the natural variables of G are p and T . So the total differential dG can also be expressed
∂G ∂G
dG = ( ) dp + ( ) dT (6.5.3)
∂p ∂T
T p
and
∂G
( ) = −S
∂T
p
which is an extraordinarily useful relationship, since one of the terms is expressible entirely in terms of measurable quantities!
∂V
( ) =V α
∂T
p
which is simply the molar volume. For a fairly incompressible substance (such as a liquid or a solid) the molar volume will be
essentially constant over a modest pressure range.
The density of gold is 19.32 g/cm3. Calculate ΔG for a 1.00 g sample of gold when the pressure on it is increased from 1.00
atm to 2.00 atm.
Solution
The change in the Gibbs function due to an isothermal change in pressure can be expressed as
6.5.1 https://chem.libretexts.org/@go/page/84327
p2
∂G
ΔG = ∫ ( ) dp
p1
∂p
T
ΔG = ∫ V dp
p1
Assuming that the molar volume is independent or pressure over the stated pressure range, ΔG becomes
ΔG = V (p2 − p1 )
So, the molar change in the Gibbs function can be calculated by substituting the relevant values.
197.0 g 1 1L 8.315 J
ΔG = ( × × ) (2.00 atm − 1.00 atm)( ) (6.5.5)
3
mol 19.32 g 1000 cm 0.08206 atm L
conversion unit
= 1.033 J (6.5.6)
Contributors
Patrick E. Fleming (Department of Chemistry and Biochemistry; California State University, East Bay)
This page titled 6.5: Pressure Dependence of Gibbs Energy is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by Patrick Fleming.
6.5.2 https://chem.libretexts.org/@go/page/84327
6.6: Temperature Dependence of A and G
In differential form, the free energy functions can be expressed as
dA = −pdV − SdT
and
dG = −V dp − SdT
and
∂G
( ) = −S
∂T p
And so, it should be fairly straightforward to determine how each changes with changing temperature:
T2 T2
∂A
ΔA = − ∫ ( ) dT = − ∫ S dT
T1 ∂T T1
V
and
T2 T2
∂G
ΔG = − ∫ ( ) dT = − ∫ S dT
T1
∂T T1
p
But the temperature dependence of the entropy needed to be known in order to evaluate the integral. A convenient work-around can
be obtained starting from the definitions of the free energy functions.
A = U −TS
and
G = H −TS
Dividing by T yields
A U
= −S
T T
and
G H
= −S
T T
and
G
⎛∂( )⎞
T H
⎜ ⎟ =−
2
∂T T
⎝ ⎠
p
Or differentiating with respect to 1/T provides a simpler form that is mathematically equivalent:
6.6.1 https://chem.libretexts.org/@go/page/84328
A
∂( )
T
( ) =U
1
∂( )
T V
and
G
⎛∂( )⎞
T
⎜ ⎟ =H
1
∂( )
⎝ T ⎠
p
Focusing on the second expression (since all of the arguments apply to the first as well), we see a system that can be integrated.
Multiplying both sides by d(1/T ) yields:
G 1
d( ) = Hd ( )
T T
and integration, assuming the enthalpy change is constant over the temperature interval yields
T2 T2
ΔG 1
∫ d( ) = ΔH ∫ d( )
T1
T T1
T
ΔGT2 ΔGT1 1 1
− = ΔH ( − ) (6.6.1)
T2 T1 T2 T1
Equation ??? is the Gibbs-Helmholtz equation and can be used to determine how ΔG changes with changing temperature. The
equivalent equation for the Helmholtz function is
ΔAT ΔAT 1 1
2 1
− = ΔU ( − ) (6.6.2)
T2 T1 T2 T1
Example 6.6.1:
Given the following data at 298 K, calculate ΔG at 500 K for the following reaction:
Compound ΔG
o
f
(kJ/mol) ΔH
f
o
(kJ.mol)
Solution
ΔH and ΔG 298 K and can be calculated fairly easily. It will be assumed that ΔH is constant over the temperature range of
298 K – 500 K.
6.6.2 https://chem.libretexts.org/@go/page/84328
ΔG500 K = −787.9 kJ
Note: ΔG became a little bit less negative at the higher temperature, which is to be expected for a reaction which is
exothermic. An increase in temperature should tend to make the reaction less favorable to the formation of products, which is
exactly what is seen in this case!
This page titled 6.6: Temperature Dependence of A and G is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated
by Patrick Fleming.
6.6.3 https://chem.libretexts.org/@go/page/84328
6.7: When Two Variables Change at Once
So far, we have derived a number of expressions and developed methods for evaluating how thermodynamic variables change as
one variable changes while holding the rest constant. But real systems are seldom this accommodating. For example, a piece of
metal (such as a railroad rail) left in the sun will undergo both an increase in temperature and an expansion due to the absorption of
energy from sunlight. So both T and V are changing at the same time! If the change in a thermodynamic variable (such as G) is
needed, contributions from both changes are required to be taken into account. We’ve already seen how to express this in terms of a
total differential.
∂G ∂G
dG = ( ) dp + ( ) dT (6.7.1)
∂p ∂T
T p
Fortunately, G (like the other thermodynamic functions U , H , S , and A ) is kind enough to be a state variable. This means that we
can consider the changes independently and then simply add the results. Another way to think of this is that the system may follow
either of two pathways to get from the initial conditions to the final conditions:
Pathway I:
1. An isothermal expansion from V to V at T followed by
1 2 1
Pathway 2:
1. An isochoric temperature increase from T to T at V followed by
1 2 1
And since G has the good sense to be a state variable, the pathway connecting the initial and final states is unimportant. We are free
to choose any path that is convenient to calculate the change.
Solution
If one considers entropy to be a function of temperature and volume, one can write the total differential of entropy as
∂S ∂S
dS = ( ) dT + ( ) dV
∂T ∂V
V T
and thus
T2 V2
∂S ∂S
ΔS = ∫ ( ) dT + ∫ ( ) dV
T1
∂T V1
∂V
V T
T1
∂T
V
T2
nCV
=∫ dT (6.7.3)
T1
T
T2
= nCV ln( ) (6.7.4)
T1
3 J 297 K
= (1.00 mol) ( ⋅ 8.314 ) ln( ) (6.7.5)
2 mol K 273 K
6.7.1 https://chem.libretexts.org/@go/page/84508
The second term is the contribution due to an isothermal expansion:
V2
∂S
ΔSV1 →V2 = ∫ ( ) dV (6.7.7)
V1
∂V
T
V1 ∂T
V
V2
nR
=∫ ( ) dV (6.7.9)
V1
V
V2
= nR ln( ) (6.7.10)
V1
J 22.0 L
= (1.00 mol) (8.314 ) ln( ) (6.7.11)
mol K 10.0 L
Thermodynamics involves many variables. But for a single component sample of matter, only two state variables are needed to
describe the system and fix all of the thermodynamic properties of the system. As such, it is conceivable that two functions can
be specified as functions of the same two variables. In general terms: z(x, y) and w(x, y).
So an important question that can be answered is, “What happens to z if w is held constant, but x is changed?” To explore this,
consider the total differential of z :
∂z ∂z
dz = ( ) dx + ( ) dy (6.7.16)
∂x ∂y
y x
but z can also be considered a function of x and w(x, y). This implies that the total differential can also be written as
∂z ∂z
dz = ( ) dx + ( ) dy (6.7.17)
∂x ∂w
w x
If we constrain the system to a change in which w remains constant, the last term will vanish since dw = 0 .
∂z ∂z ∂z
( ) dx + ( ) dy = ( ) dx (6.7.18)
∂x ∂y ∂x
y x w
but also, since w is a function x and y , the total differential for w can be written
∂w ∂w
dw = ( ) dx + ( ) dy
∂x ∂y
y x
6.7.2 https://chem.libretexts.org/@go/page/84508
And it too must be zero for a process in which w is held constant.
∂w ∂w
0 =( ) dx + ( ) dy
∂x ∂y
y x
∂z ∂z ∂w ∂y ∂z
( ) dx + ( ) [−( ) ( ) dx] = ( ) dx (6.7.19)
∂x y
∂y x
∂x y
∂w x
∂x w
which simplifies to
∂z ∂z ∂w ∂z
( ) dx − ( ) ( ) dx = ( ) dx
∂x ∂w ∂x ∂x
y x y w
or
∂z ∂z ∂z ∂w
( ) =( ) +( ) ( ) (6.7.20)
∂x ∂x ∂w ∂x
y w x y
As with partial derivative transformation types I and II, this result can be achieved in a formal, albeit less mathematically
rigorous method.
Consider z(x, w). This allows us to write the total differential for z :
∂z ∂z
dz = ( ) dx + ( ) dw
∂x ∂w
w x
noting that dx/dx = 1 and converting the other ratios to partial derivatives yields
∂z ∂z ∂z ∂w
( ) =( ) +( ) ( ) (6.7.21)
∂x ∂x ∂w ∂x
y w x y
which agrees with the previous result (Equation 6.7.20)! Again, the method is not mathematically rigorous, but it works so
long as w, x, y , and z are state functions and the total differentials dw, dx, dy , and dz are exact.
Contributors
Patrick E. Fleming (Department of Chemistry and Biochemistry; California State University, East Bay)
This page titled 6.7: When Two Variables Change at Once is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated
by Patrick Fleming.
6.7.3 https://chem.libretexts.org/@go/page/84508
6.8: The Difference between Cp and Cv
Constant volume and constant pressure heat capacities are very important in the calculation of many changes. The ratio
C /C
p V = γ appears in many expressions as well (such as the relationship between pressure and volume along an adiabatic
expansion.) It would be useful to derive an expression for the difference C – C as well. As it turns out, this difference is
p V
and
∂U
CV = ( )
∂T
V
H = U + pV
dH = dU + pdV + V dp
The last term is kind enough to vanish (since dp = 0 at constant pressure). After converting the remaining terms to partial
derivatives:
∂H ∂U ∂V
( ) =( ) + p( ) (6.8.1)
∂T p
∂T p
∂T p
and
∂V
( ) =V α
∂T
p
In order to evaluate the partial derivative above, first consider U (V , T ) . Then the total differential du can be expressed
∂U ∂U
du = ( ) dV − ( ) dT
∂V T
∂T V
Dividing by dT and constraining to constant p will generate the partial derivative we wish to evaluate:
6.8.1 https://chem.libretexts.org/@go/page/84509
dU ∣ ∂U dV ∣ ∂U dT ∣
∣ =( ) ∣ +( ) ∣
dT ∣ p
∂V dT ∣ p
∂T dT ∣ p
T V
The last term will become unity, so after converting to partial derivatives, we see that
∂U ∂U ∂V ∂U
( ) =( ) ( ) +( ) (6.8.3)
∂T ∂V ∂T ∂T
p T p V
(This, incidentally, is an example of partial derivative transformation type III.) Now we are getting somewhere!
∂U
( ) = CV
∂T
V
and
∂V
( ) =V α
∂T
p
we are almost home free! Fortunately, that is an easy expression to derive. Begin with the combined expression of the first and
second laws:
d = T dS − pdV
and since
∂p α
( ) =
∂T κT
V
then
∂U α
( ) =T −p
∂V κT
T
Now, substituting this into the expression into Equation 6.8.3 to get
6.8.2 https://chem.libretexts.org/@go/page/84509
∂U α
( ) = [T − p] V α + CV
∂T κT
p
2
TV α
= − pV α + CV
κT
The pV α terms will cancel. And subtracting C from both sides gives the desired result:
V
2
TV α
Cp − CV = (6.8.5)
κT
And this is a completely general result since the only assumptions made were those that allowed us to use the combined first and
second laws in the form
dU = T dS– pdV .
That means that this expression can be applied to any substance whether gas, liquid, animal, vegetable, or mineral. But what is the
result for an ideal gas?
Since we know that for an ideal gas
1
α =
T
and
1
κT =
p
pV
=
T
=R
Example 6.8.1
Derive the expression for the difference between C and C by beginning with the definition of H , differentiating, dividing by
p V
dV (to generate the partial derivative definition of C ). In this approach, you will need to find expressions for
V
∂H
( )
∂T
V
and
∂U
( )
∂p
T
6.8.3 https://chem.libretexts.org/@go/page/84509
Solution
Begin with the definition of enthalpy.
H = U + pV
dH = dU + pdV + V dp
Now, divide by dV and constrain to constant T (as described in the instructions) to generate the partial derivative definition of
CV
dH ∣ dU ∣ dV ∣ dp ∣
∣ = ∣ +p ∣ +V ∣
dT ∣ V
dT ∣ V
dT ∣ V
dT ∣V
dH dU dp
( ) =( ) +V ( ) (6.8.6)
dT dT dT
V V V
This can be derived from the total differential for H (p, T ) by dividing by dT and constraining to constant V .
dH dH
dH = ( ) dp + ( ) dT
dp dT
T p
dH dH dp dH
( ) =( ) ( ) +( ) (6.8.7)
dT dp dT dT
V T V p
This again is an example of partial derivative transformation type III. To continue, we need an expression for
dH
( ) .
dp
T
This can be quickly generated by considering the total differential of H (p, S), its natural variables:
dH = T dS + V dp
dH dS
( ) = T( ) +V (6.8.8)
dp dp
T T
Now, substitute this back into the expression for (Equation 6.8.7):
dH dV dp dH
( ) = [−T ( ) +V ] ( ) +( )
dT dT dT dT
V p V p
6.8.4 https://chem.libretexts.org/@go/page/84509
dH dV dp dp dH
( ) = −T ( ) ( ) +V ( ) +( )
dT dT dT dT dT
V p V V p
dH
This can now substituted for the right-hand side of the initial expression for ( ) back into Equation 6.8.6:
dT
V
dV dp dp dH dU dp
−T ( ) ( ) + V ( ) +( ) =( ) + V ( ) (6.8.9)
dT dT dT dT dT dT
p V V p V V
Several terms cancel one another. Equation 6.8.9 can then be rearranged to yield
dH dU dV dp
( ) −( ) = T( ) ( )
dT dT dT dT
p V p V
or
2
TV α
Cp − CV =
κT
This page titled 6.8: The Difference between Cp and Cv is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated
by Patrick Fleming.
6.8.5 https://chem.libretexts.org/@go/page/84509
6.E: Putting the Second Law to Work (Exercises)
Q6.1
Using Table T1, calculate the standard reaction Gibbs functions (\(\Delta G^o\)) for the following reactions at 298 K.
a. C H C H OH (l) + 3O (g) → 2C O (g) + 3H
3 2 2 2 2 O(l)
Q6.2
Estimate \(\Delta G\) at 1000 K from its value at 298 K for the reaction
Q6.3
The standard Gibbs function for formation (ΔG ) of P bO (s) is -217.4 kJ/mol at 298 K. Assuming
o
f 2 O2 is an ideal gas, find the
standard Helmholtz function for formation (ΔA for P bO at 298K. o
f 2
Q6.4
Calculate the entropy change for 1.00 mol of an ideal monatomic gas (CV = 3/2 R) undergoing an expansion and simultaneous
temperature increase from 10.0 L at 298 K to 205.0 L at 455 K.
Q6.5
Consider a gas that obeys the equation of state
nRT
p = (6.E.2)
V − nb
Q6.6
Show that
∂Cp
( ) =0 (6.E.3)
∂p
T
Q6.7
Derive the thermodynamic equation of state
∂H
( ) = V (1 − T α) (6.E.4)
∂p
T
Q6.8
Derive the thermodynamic equation of state
∂U α
( ) =T −p (6.E.5)
∂V κT
T
6.E.1 https://chem.libretexts.org/@go/page/84511
Q6.9
The “Joule Coefficient” is defined by
∂T
μJ = ( ) (6.E.6)
∂V U
Show that
1 Tα
μJ = (p − ) (6.E.7)
CV κT
Q6.10
Derive expressions for the pressure derivatives
∂X
( ) (6.E.8)
∂p
T
∂H
whereX is U , H , A , G, and S at constant temperature in terms of measurable properties. (The derivation of ( ) was done in
∂p
T
problem Q6.7).
Evaluate the expressions for
∂S
( )
∂p
T
∂H
( )
∂p
T
∂U
( )
∂p
T
Q6.11
Derive expressions for the volume derivatives
∂X
( ) (6.E.9)
∂V
T
∂U
where X is U , H , A , G, and S at constant temperature in terms of measurable properties. (The derivation of ( ) was done in
∂V T
problem Q8.8.)
Evaluate the expressions for
∂X
( )
∂V
T
∂X
( )
∂V
T
Q6.12
Evaluate the difference between C and C for a gas that obeys the equation of state
p V
nRT
p = (6.E.10)
V − nb
6.E.2 https://chem.libretexts.org/@go/page/84511
Q6.13
The adiabatic compressibility (k ) is defined by
S
1 ∂V
κS = ( ) (6.E.11)
V ∂p S
This page titled 6.E: Putting the Second Law to Work (Exercises) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by Patrick Fleming.
6.E.3 https://chem.libretexts.org/@go/page/84511
6.S: Putting the Second Law to Work (Summary)
Learning Objectives
After mastering the material presented in this chapter, one will be able to:
1. Define the free energy functions A and G, and relate changes in these functions to the spontaneity of a given process and
constant volume and pressure respectively.
2. Use the definitions of entropy and reversible work of expansion to write an equation that combines the first and second laws of
thermodynamics.
3. Utilize the combined first and second law relationship to derive Maxwell Relations stemming from the definitions of U , H , A ,
and G.
4. Utilize the Maxwell Relations to derive expressions that govern changes in thermodynamic variable as systems move along
specified pathways (such as constant temperature, pressure, volume, or adiabatic pathways.)
5. Derive and utilize an expression describing the volume dependence of A .
6. Derive and utilize an expression describing the pressure dependence of G.
7. Derive and utilize expressions that describe the temperature, dependence of A and G.
8. Derive an expression for, and evaluate the difference between C and C for any substance, in terms of T , V , α , and κ .
p V T
This page titled 6.S: Putting the Second Law to Work (Summary) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by Patrick Fleming.
6.S.1 https://chem.libretexts.org/@go/page/84510
CHAPTER OVERVIEW
This page titled 7: Mixtures and Solutions is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
1
7.1: Thermodynamics of Mixing
A natural place to begin a discussion of mixtures is to consider a mixture of two gases. Consider samples of the two gases filling
two partitions in a single container, both at the same pressure, temperature, having volumes V and V .
A B
After being allowed to mix isothermally, the partial pressures of the two gases will drop by a factor of 2 (although the total pressure
will still be the original value) and the volumes occupied by the two gases will double.
Enthalpy of Mixing
Assuming ideal behavior, so that interactions between individual gas molecules are unimportant, it is fairly easy to calculate \(Delta
H) for each gas, as it is simply an isothermal expansion. The total enthalpy of mixing is then given by
And since the enthalpy change for an isothermal expansion of an ideal gas is zero,
ΔHmix = 0
Entropy of Mixing
The entropy change induced due to isothermal mixing (assuming again no interactions between the molecules in the gas mixture) is
again going to be the sum of the contributions from isothermal expansions of the two gases. Fortunately, entropy changes for
isothermal expansions are easy to calculate for ideal gases.
V2
ΔS = nR ln( )
V1
If we use the initial volumes VA and VB for the initial volumes of gases A and B, the total volume after mixing is V A + VB , and the
total entropy change is
VA + VB VA + VB
ΔSmix = nA R ln( ) + nA R ln( )
VA VB
where χ is the mole fraction of A after mixing, and that n can be expresses as the product of χ and the total number of moles,
A A A
7.1.1 https://chem.libretexts.org/@go/page/84330
ΔSmix = ntot R [−χA ln(χA ) − χB ln(χB )]
It should be noted that because the mole fraction is always between 0 and 1, that ln(χ ) < 0 . As such, the entropy change for a
B
system undergoing isothermal mixing is always positive, as one might expect (since mixing will make the system less ordered).
The entropy change for a system undergoing isothermal mixing is always positive.
and so it follows from above that for the isothermal mixing of two gases at constant total pressure
The relationships describing the isothermal mixing of two ideal gases A and B is summarized in the graph below.
true for gases. But for many combinations of liquids or solids, the strong intermolecular forces may make mixing unfavorable (for
example in the case of vegetable oil and water). Also, these interactions may make the volume non-additive as well (as in the case
of ethanol and water).
This page titled 7.1: Thermodynamics of Mixing is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
7.1.2 https://chem.libretexts.org/@go/page/84330
7.2: Partial Molar Volume
The partial molar volume of compound A in a mixture of A and B can be defined as
∂V
VA = ( )
∂nA
p,T ,nB
Using this definition, a change in volume for the mixture can be described using the total differential of V :
∂V ∂V
dV = ( ) dnA + ( ) dnB
∂nA ∂nB
p,T ,nB p,T ,nA
or
dV = Va dnA + Vb dnB
V =∫ Va dnA + ∫ Vb dnB
0 0
V = Va nA + Vb nB
This result is important as it demonstrates an important quality of partial molar quantities. Specifically, if ξi represents the partial
molar property X for component i of a mixture, The total property X for the mixture is given by
X = ∑ ξi ni
It should be noted that while the volume of a substance is never negative, the partial molar volume can be. An example of this
appears in the dissolution of a strong electrolyte in water. Because the water molecules in the solvation sphere of the ions are
physically closer together than they are in bulk pure water, there is a volume decrease when the electrolyte dissolves. This is easily
observable at high concentrations where a larger fraction of the water in the sample is tied up in solvation of the ions.
This page titled 7.2: Partial Molar Volume is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
7.2.1 https://chem.libretexts.org/@go/page/84331
7.3: Chemical Potential
In much the same fashion as the partial molar volume is defined, the partial molar Gibbs function is defined for compound i in a
mixture:
∂G
μi = ( ) (7.3.1)
∂ni
p,T , nj ≠i
This particular partial molar function is of particular importance, and is called the chemical potential. The chemical potential tells
how the Gibbs function will change as the composition of the mixture changes. And since systems tend to seek a minimum
aggregate Gibbs function, the chemical potential will point to the direction the system can move in order to reduce the total Gibbs
function. In general, the total change in the Gibbs function (dG) can be calculated from
∂G ∂G ∂G
dG = ( ) dp + ( ) dT + ∑ ( ) dni
∂p ∂T ∂ni
T ,ni p,ni i T , nj ≠i
Or, by substituting the definition for the chemical potential, and evaluating the pressure and temperature derivatives as was done in
Chapter 6:
dG = V dp − SdT + ∑ μi dni
But as it turns out, the chemical potential can be defined as the partial molar derivative any of the four major thermodynamic
functions U , H , A , or G:
Table 7.3.1 : Chemical potential can be defined as the partial molar derivative any of the four major thermodynamic functions
∂U
dU = T dS − pdV + ∑ μi dni μi = ( )
i ∂ni
S,V ,nj ≠i
∂H
dH = T dS − V dT + ∑ μi dni μi = ( )
i ∂ni
S,p,nj ≠i
∂A
dA = −pdV − T dS + ∑ μi dni μi = ( )
i ∂ni
V ,T ,nj ≠i
∂G
dG = V dp − SdT + ∑ μi dni μi = ( )
i ∂ni
p,T ,nj ≠i
The last definition, in which the chemical potential is defined as the partial molar Gibbs function is the most commonly used, and
perhaps the most useful (Equation 7.3.1). As the partial most Gibbs function, it is easy to show that
dμ = V dp − SdT
where V is the molar volume, and S is the molar entropy. Using this expression, it is easy to show that
∂μ
( ) =V
∂p
T
∫ dμ = ∫ V dp (7.3.2)
o o
μ p
So that for a substance for which the molar volume is fairly independent of pressure at constant temperature (i. e., κT is very
small), therefore Equation 7.3.2 becomes
μ p
∫ dμ = V ∫ dp
o o
μ p
o o
μ−μ = V (p − p )
or
7.3.1 https://chem.libretexts.org/@go/page/84332
o o
μ =μ + V (p − p )
Where p is a reference pressure (generally the standard pressure of 1 atm) and μ is the chemical potential at the standard
o o
pressure. If the substance is highly compressible (such as a gas) the pressure dependence of the molar volume is needed to
complete the integral. If the substance is an ideal gas
RT
V =
p
or
p
o
μ =μ + RT ln( )
o
p
This page titled 7.3: Chemical Potential is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
7.3.2 https://chem.libretexts.org/@go/page/84332
7.4: The Gibbs-Duhem Equation
For a system at equilibrium, the Gibbs-Duhem equation must hold:
∑ ni dμi = 0 (7.4.1)
This relationship places a compositional constraint upon any changes in the chemical potential in a mixture at constant temperature
and pressure for a given composition. This result is easily derived when one considers that μ represents the partial molar Gibbs
i
Gtot = ∑ ni μi
i i
dG = V dp − sdT + ∑ μi dni
i i i
Substituting Equation 7.4.3 into 7.4.2 results in the Gibbs-Duhem equation (Equation 7.4.1). This expression relates how the
chemical potential can change for a given composition while the system maintains equilibrium. So for a binary system, consisting
of components A and B (the two most often studied compounds in all of chemistry)
nA
dμB = − dμA
nB
This page titled 7.4: The Gibbs-Duhem Equation is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
7.4.1 https://chem.libretexts.org/@go/page/84333
7.5: Non-ideality in Gases - Fugacity
The relationship for chemical potential
o
p
μ =μ + RT ln( )
o
p
was derived assuming ideal gas behavior. But for real gases that deviate widely from ideal behavior, the expression has only
limited applicability. In order to use the simple expression on real gases, a “fudge” factor is introduced called fugacity. Using
fugacity instead of pressure, the chemical potential expression becomes
o
f
μ =μ + RT ln( )
o
f
where f is the fugacity. Fugacity is related to pressure, but contains all of the deviations from ideality within it. To see how it is
related to pressure, consider that a change in chemical potential for a single component system can be expressed as
dμ − V dp − SdT
and so
∂μ
( ) =V (7.5.1)
∂p
T
Differentiating the expression for chemical potential above with respect to pressure at constant volume results in
∂μ ∂ o
f
( ) ={ [μ + RT ln( )]}
o
∂p ∂p f
T
which simplifies to
∂μ ∂ ln(f )
( ) = RT [ ] =V
∂p ∂p
T T
where Z is the compression factor as discussed previously. Now, we can use the expression above to obtain the fugacity
coefficient γ, as defined by
f = γp
ln f = ln γ + ln p
or
ln γ = ln f − ln p
Z 1
=∫ ( − ) dp
∂p ∂p
T
7.5.1 https://chem.libretexts.org/@go/page/84334
If the gas behaves ideally, γ = 1 . In general, this will be the limiting value as p → 0 since all gases behave ideal as the pressure
approaches 0. The advantage to using the fugacity in this manner is that it allows one to use the expression
o
f
μ =μ + RT ln( )
o
f
to calculate the chemical potential, insuring that Equation 7.5.1 holds even for gases that deviate from ideal behavior!
This page titled 7.5: Non-ideality in Gases - Fugacity is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
7.5.2 https://chem.libretexts.org/@go/page/84334
7.6: Colligative Properties
Colligative properties are important properties of solutions as they describe how the properties of the solvent will change as solute
(or solutes) is (are) added. Before discussing these important properties, let us first review some definitions.
Solution – a homogeneous mixture.
Solvent – The component of a solution with the largest mole fraction
Solute – Any component of a solution that is not the solvent.
Solutions can exist in solid (alloys of metals are an example of solid-phase solutions), liquid, or gaseous (aerosols are examples of
gas-phase solutions) forms. For the most part, this discussion will focus on liquid-phase solutions.
μsolid ≤ μliquid
As such, the freezing point of the solvent in a solution will be affected by anything that changes the chemical potential of the
solvent. As it turns out, the chemical potential of the solvent is reduced by the presence of a solute.
In a mixture, the chemical potential of component A can be calculated by
o
μA = μ + RT ln χA (7.6.1)
A
And because χ is always less than (or equal to) 1, the chemical potential is always reduced by the addition of another component.
A
μA,solid = μA,liquid
where the chemical potential of the liquid is given by Equation 7.6.1, which rearrangement to
o
μA − μ
A
= ln χA
RT
To evaluate the temperature dependence of the chemical potential, it is useful to consider the temperature derivative at constant
pressure.
o
∂ μA − μ ∂ ln χA
A
[ ( )] =( )
∂T RT ∂T
p p
o o
μA − μ 1 ∂μA ∂μ ∂ ln χA
A A
− + [( ) −( ) ] =( ) (7.6.2)
2
RT RT ∂T ∂T ∂T
p p p
Recalling that
μ = H = TS
and
∂μ
( ) = −S
∂T
p
And noting that in the case of the solvent freezing, H is the enthalpy of the pure solvent in solid form, and H is the enthalpy of
o
A
A
7.6.1 https://chem.libretexts.org/@go/page/84335
Equation 7.6.3 then becomes
o o
ΔHf us −SA + S −SA + S ∂ ln χA
A A
− + =( )
2
RT RT RT ∂T
p
or
ΔHf us ∂ ln χA
=( )
2
RT ∂T
p
where ΔT is the difference between the freezing temperature of the pure solvent and that of the solvent in the solution. Also, for
small deviations from the pure freezing point, T T can be replaced by the approximate value (T ) . So the Equation 7.6.5
o o 2
becomes
ΔHf us
− ΔT = ln χA (7.6.6)
o 2
R(T )
Further, for dilute solutions, for which χ , the mole fraction of the solvent is very nearly 1, then
A
ln χA ≈ −(1 − χA ) = −χB
where χ is the mole fraction of the solute. After a small bit of rearrangement, this results in an expression for freezing point
B
depression of
o 2
R(T )
ΔT = ( ) χB
ΔHf us
o 2
R(T )
= Kf
ΔHf us
freezing point depression property is independent of the solute and is a property based solely on the nature of the solvent. Further,
since χ was introduced as (1 − χ ) , it represents the sum of the mole fractions of all solutes present in the solution.
B A
It is important to keep in mind that for a real solution, freezing of the solvent changes the composition of the solution by decreasing
the mole fraction of the solvent and increasing that of the solute. As such, the magnitude of ΔT will change as the freezing process
continually removes solvent from the liquid phase of the solution.
7.6.2 https://chem.libretexts.org/@go/page/84335
vapor phase. As sch, the temperature must be increased to increase the chemical potential of the solvent in the liquid solution until
it is equal to that of the vapor-phase solvent. The increase in the boiling point can be expressed as
ΔT = Kb χB
where
o 2
R(T )
= Kb
ΔHvap
is called the ebullioscopic constant and, like the cryoscopic constant, is a property of the solvent that is independent of the solute
or solutes. A very elegant derivation of the form of the models for freezing point depression and boiling point elevation has been
shared by F. E. Schubert (Schubert, 1983).
Cryoscopic and ebullioscopic constants are generally tabulated using molality as the unit of solute concentration rather than mole
fraction. In this form, the equation for calculating the magnitude of the freezing point decrease or the boiling point increase is
ΔT = Kf m
or
ΔT = Kb m
where m is the concentration of the solute in moles per kg of solvent. Some values of K and K are shown in the table below.
f b
Example 7.6.1:
The boiling point of a solution of 3.00 g of an unknown compound in 25.0 g of CCl4 raises the boiling point to 81.5 °C. What
is the molar mass of the compound?
Solution
The approach here is to find the number of moles of solute in the solution. First, find the concentration of the solution:
m = 0.936 mol/kg
Using the number of kg of solvent, one finds the number for moles of solute:
7.6.3 https://chem.libretexts.org/@go/page/84335
In order to establish equilibrium between the solvent in the solution and the solvent in the vapor phase above the solution, the
chemical potentials of the two phases must be equal.
μvapor = μsolvent
If the solute is not volatile, the vapor will be pure, so (assuming ideal behavior)
′
p
o o
μvap + RT ln =μ + RT ln χA (7.6.7)
o A
p
Where p is the vapor pressure of the solvent over the solution. Similarly, for the pure solvent in equilibrium with its vapor
′
pA
o o
μ = μvap + RT ln (7.6.8)
A o
p
Subtracting RT ln(P A /P
o
) from both side produces
′
p pA
RT ln − RT ln = RT ln χA
o o
p p
which rearranges to
′
p
RT ln = RT ln χA
pA
or
′
p = χA pA (7.6.9)
This last result is Raoult’s Law. A more formal derivation would use the fugacities of the vapor phases, but would look essentially
the same. Also, as in the case of freezing point depression and boiling point elevations, this derivation did not rely on the nature of
the solute! However, unlike freezing point depression and boiling point elevation, this derivation did not rely on the solute being
dilute, so the result should apply the entire range of concentrations of the solution.
Example 7.6.2:
Consider a mixture of two volatile liquids A and B. The vapor pressure of pure A is 150 Torr at some temperature, and that of
pure B is 300 Torr at the same temperature. What is the total vapor pressure above a mixture of these compounds with the mole
fraction of B of 0.600. What is the mole fraction of B in the vapor that is in equilibrium with the liquid mixture?
Solution
Using Raoult’s Law (Equation 7.6.9)
7.6.4 https://chem.libretexts.org/@go/page/84335
ptot = pA + pB = 240 T orr
To get the mole fractions in the gas phase, one can use Dalton’s Law of partial pressures.
pA 60.0 T orr
χA = = = 0.250
ptot 240 T orr
pB 180.0 T orr
χB = = = 0.750
ptot 240 T orr
And, of course, it is also useful to note that the sum of the mole fractions is 1 (as it must be!)
χA + χB = 1
Osmotic Pressure
Osmosis is a process by which solvent can pass through a semi-permeable membrane (a membrane through which solvent can
pass, but not solute) from an area of low solute concentration to a region of high solute concentration. The osmotic pressure is the
pressure that when exerted on the region of high solute concentration will halt the process of osmosis.
The nature of osmosis and the magnitude of the osmotic pressure can be understood by examining the chemical potential of a pure
solvent and that of the solvent in a solution. The chemical potential of the solvent in the solution (before any extra pressure is
applied) is given by
o
μA = μ + RT ln χA
A
And since xA < 1, the chemical potential is of the solvent in a solution is always lower than that of the pure solvent. So, to prevent
osmosis from occurring, something needs to be done to raise the chemical potential of the solvent in the solution. This can be
accomplished by applying pressure to the solution. Specifically, the process of osmosis will stop when the chemical potential
solvent in the solution is increased to the point of being equal to that of the pure solvent. The criterion, therefore, for osmosis to
cease is
o
μ (p) = μA (χb , +π)
A
To solve the problem to determine the magnitude of p, the pressure dependence of the chemical potential is needed in addition to
understanding the effect the solute has on lowering the chemical potential of the solvent in the solution. The magnitude, therefore,
of the increase in chemical potential due to the application of excess pressure p must be equal to the magnitude of the reduction of
chemical potential by the reduced mole fraction of the solvent in the solution. We already know that the chemical potential of the
solvent in the solution is reduced by an amount given by
o
μ − μA = RT ln χA
A
And the increase in chemical potential due to the application of excess pressure is given by
π
∂μ
μ(p + π) = μ(p) + ∫ ( ) dp
p
∂p
T
7.6.5 https://chem.libretexts.org/@go/page/84335
∂μ
( ) =V
∂p
T
where V is the molar volume of the substance. Combining these expressions results in
p+π
−RT ln χA = ∫ V dp
p
If the molar volume of the solvent is independent of pressure (has a very small value of κ – which is the case for most liquids) the
T
ln χA ≈ −(1 − χA ) = −χB
χB RT = V π
Or after rearrangement
χB RT
π
V
again, where V is the molar volume of the solvent. And finally, since χ /V is the concentration of the solute B for cases where
B
nB ≪ n . This allows one to write a simplified version of the expression which can be used in the case of very dilute solutions
A
π = [B]RT
When a pressure exceeding the osmotic pressure π is applied to the solution, the chemical potential of the solvent in the solution
can be made to exceed that of the pure solvent on the other side of the membrane, causing reverse osmosis to occur. This is a very
effective method, for example, for recovering pure water from a mixture such as a salt/water solution.
This page titled 7.6: Colligative Properties is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
7.6.6 https://chem.libretexts.org/@go/page/84335
7.7: Solubility
The maximum solubility of a solute can be determined using the same methods we have used to describe colligative properties. The
chemical potential of the solute in a liquid solution can be expressed
o
μB (solution) = μ (liquid) + RT ln χB
B
If this chemical potential is lower than that of a pure solid solute, the solute will dissolve into the liquid solvent (in order to achieve
a lower chemical potential!) So the point of saturation is reached when the chemical potential of the solute in the solution is equal
to that of the pure solid solute.
o o
μ (solid) = μ (liquid) + RT ln χB
B B
Since the mole fraction at saturation is of interest, we can solve for ln(χ B) .
o o
μ (solid) = μ (liquid)
B B
ln χB =
RT
The difference in the chemical potentials is the molar Gibbs function for the phase change of fusion. So this can be rewritten
o
−ΔG
f us
ln χB =
RT
It would be convenient if the solubility could be expressed in terms of the enthalpy of fusion for the solute rather than the Gibbs
function change. Fortunately, the Gibbs-Helmholtz equation gives us a means of making this change. Noting that
ΔG
⎛ ⎞
∂( )
⎜ T ⎟ ΔH
⎜ ⎟ =
⎜ ⎟ 2
∂T T
⎝ ⎠
p
Differentiation of the above expression for ln(χ B) with respect to T at constant p yields
∂ ln χB 1 ΔHf us
( ) =
2
∂T R T
p
Separating the variables puts this into an integrable form that can be used to see how solubility will vary with temperature:
ln χB T
1 ΔHf us dT
∫ d ln χB = ∫
2
0
R Tf T
So if the enthalpy of fusion is constant over the temperature range of T to the temperature of interest,
f
ΔHf us 1 1
ln χB = ( − )
R Tf T
And χ will give the mole fraction of the solute in a saturated solution at the temperature
B T . The value depends on both the
enthalpy of fusion, and the normal melting point of the solute.
This page titled 7.7: Solubility is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick Fleming.
7.7.1 https://chem.libretexts.org/@go/page/84528
7.8: Non-ideality in Solutions - Activity
The bulk of the discussion in this chapter dealt with ideal solutions. However, real solutions will deviate from this kind of behavior.
So much as in the case of gases, where fugacity was introduced to allow us to use the ideal models, activity is used to allow for the
deviation of real solutes from limiting ideal behavior. The activity of a solute is related to its concentration by
mB
aB = γ
o
m
where γ is the activity coefficient, m is the molaliy of the solute, and m is unit molality. The activity coefficient is unitless in
B
o
this definition, and so the activity itself is also unitless. Furthermore, the activity coefficient approaches unity as the molality of the
solute approaches zero, insuring that dilute solutions behave ideally. The use of activity to describe the solute allows us to use the
simple model for chemical potential by inserting the activity of a solute in place of its mole fraction:
o
μB = μ + RT ln aB
B
The problem that then remains is the measurement of the activity coefficients themselves, which may depend on temperature,
pressure, and even concentration.
the chemical potential of the cation can be denoted μ and that of the anion as μ . For a solution, the total molar Gibbs function of
+ −
G = μ+ + μ−
where
∗
μ =μ + RT ln a
where \(\mu^*\) denotes the chemical potential of an ideal solution, and a is the activity of the solute. Substituting his into the
above relationship yields
∗ ∗
G = μ+ + RT ln a+ + μ− + RT ln a−
ai = γi mi
The expression for the total molar Gibbs function of the solutes becomes
∗ ∗
G = μ+ + RT ln γ+ m+ + μ− + RT ln γ− m−
where all of the deviation from ideal behavior comes from the last term. Unfortunately, it impossible to experimentally deconvolute
the term into the specific contributions of the two ions. So instead, we use a geometric average to define the mean activity
coefficient, γ .
±
−−− −
γ± = √γ+ γ−
7.8.1 https://chem.libretexts.org/@go/page/84529
Debeye-Hückel Law
In 1923, Debeye and Hückel (Debye & Hückel, 1923) suggested a means of calculating the mean activity coefficients from
experimental data. Briefly, they suggest that
6
1.824 × 10 –
log10 γ± = | z+ + z− | √I
3/2
(ϵT )
where ϵ is the dielectric constant of the solvent, T is the temperature in K, z and z are the charges on the ions, and I is the ionic
+ −
o
For a solution in water at 25 C,
This page titled 7.8: Non-ideality in Solutions - Activity is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated
by Patrick Fleming.
7.8.2 https://chem.libretexts.org/@go/page/84529
7.E: Mixtures and Solutions (Exercises)
Q7.1
The compression factor (Z ) for O2 at 200 K is measured to have the following values:
p (atm) Z
1.000 0.9970
4.000 0.9880
7.000 0.9788
10.000 0.9700
Using numerical integration, calculate the fugacity constant for O2 at 200 K from these data.
Q7.2
The normal boiling point of ethanol is 78.4 oC. Its enthalpy of vaporization is 38.6 kJ/mol. Estimate the vapor pressure of ethanol at
24.4 oC.
Q7.3
When 20.0 grams of an unknown nonelectrolyte compound are dissolved in 500.0 grams of benzene, the freezing point of the
resulting solution is 3.77 °C. The freezing point of pure benzene is 5.444 °C and the cryoscopic constant (K ) for benzene is 5.12
f
Q7.4
Consider a mixture of two volatile liquids, A and B. The vapor pressure of pure liquid A is 324.3 Torr and that of pure liquid B is
502.3 Torr. What is the total vapor pressure over a mixture of the two liquids for which xB = 0.675?
Q7.5
Consider the following expression for osmotic pressure
πV = χB RT (7.E.1)
where π is the osmotic pressure, V is the molar volume of the solvent, χB is the mole fraction of the solute, R is the gas law
constant, and T is the temperature (in Kelvin).
The molar volume of a particular solvent is 0.0180 L/mol. 0.200 g of a solute (B) is dissolved in 1.00 mol of the solvent. The
osmotic pressure of the solvent is then measured to be 0.640 atm at 298 K. Calculate the molar mass of the solute.
Q7.6
At 300 K, the vapor pressure of HCl(g) over a solution of HC l in GeC l4 are summarized in the following table. Calculate the
Henry’s Law constant for HCl based on these data.
0.005 32.0
0.012 76.9
0.019 121.8
Q7.7
Consider the mixing of 1.00 mol of hexane (C6H12) with 1.00 mole of benzene (C6H6). Calculate \(\Delta H\), \(\Delta S\), and \
(\Delta G\) of mixing, of the mixing occurs ideally at 298 K.
7.E.1 https://chem.libretexts.org/@go/page/84526
Contributors and Attributions
Patrick E. Fleming (Department of Chemistry and Biochemistry; California State University, East Bay)
This page titled 7.E: Mixtures and Solutions (Exercises) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated
by Patrick Fleming.
7.E.2 https://chem.libretexts.org/@go/page/84526
7.S: Mixtures and Solutions (Summary)
Learning Objectives
After mastering the material in this chapter, one will be able to
1. Describe the thermodynamics of mixing and calculate \(\Delta H\), \(\Delta S\), and \(\Delta G\) or mixing for an ideal solution.
2. Define chemical potential, and calculate its value as a function of pressure and composition.
3. Derive expressions for the colligative properties and perform calculations using the relationships.
4. Estimate the maximum solubility of a solute in a solvent based on the concept equality of chemical potential at saturation.
5. Define fugacity and activity.
6. Calculate the mean activity coefficients of ions in solution based on the ionic strength of the solution.
References
1. Debye, P., & Hückel, E. (1923). Zur Theorie der Electrolyte. Physikalische Zeitschrift, 24, 185-206.
2. Schubert, F. E. (1983). Depression of Freezing Point and Elevation of Boiling Point. Journal of Chemical Education, 60(1), 88.
doi:10.1021/ed060p87.2
This page titled 7.S: Mixtures and Solutions (Summary) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated
by Patrick Fleming.
7.S.1 https://chem.libretexts.org/@go/page/84527
CHAPTER OVERVIEW
8: Phase Equilibrium
This page titled 8: Phase Equilibrium is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick Fleming.
1
8.1: Prelude to Phase Equilibrium
From the very elementary stages of our journey to describe the physical nature of matter, we learn to classify mater into three (or
more) phases: solid, liquid, and gas. This is a fairly easy classification system that can be based on such simple ideas as shape and
volume.
As we have progressed, we have seen that solids and liquids are not completely incompressible as they may have non-zero values
of κ . And we learn that there are a number of finer points to describing the nature of the phases about which we all learn in grade
T
school. In this chapter, we will employ some of the tools of thermodynamics to explore the nature of phase boundaries and see
what we can conclude about them.
This page titled 8.1: Prelude to Phase Equilibrium is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
8.1.1 https://chem.libretexts.org/@go/page/91111
8.2: Single Component Phase Diagrams
The stability of phases can be predicted by the chemical potential, in that the most stable form of the substance will have the
minimum chemical potential at the given temperature and pressure. This can be summarized in a phase diagram like the one shown
below.
In this diagram, the phase boundaries can be determined by measuring the rate of cooling at constant temperature. A typical cooling
curve is shown below. The temperature will decrease over time as a sample is allowed to cool. When κ the substance undergoes a
T
phase change, say from liquid to solid, the temperature will stop changing while heat is extracted due to the phase change. The
temperature at which the halt occurs provides one point on the boundary at the temperature of the halt and the pressure at which the
cooling curve was measured.
The same data can be obtained by heating the system using a technique such as scanning calorimetry. In this experiment, heat is
supplied to a sample at a constant rate, and the temperature of the sample is measured, with breaks occurring at the phase change
temperatures.
This page titled 8.2: Single Component Phase Diagrams is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated
by Patrick Fleming.
8.2.1 https://chem.libretexts.org/@go/page/84337
8.3: Criterion for Phase Equilibrium
The thermodynamic criterion for phase equilibrium is simple. It is based upon the chemical potentials of the components in a
system. For simplicity, consider a system with only one component. For the overall system to be in equilibrium, the chemical
potential of the compound in each phase present must be the same. Otherwise, there will be some mass migration from one phase to
another, decreasing the total chemical potential of the phase from which material is being removed, and increasing the total
chemical potential of the phase into which the material is being deposited. So for each pair of phases present (α and β) the
following must be true:
μα = μβ
∑ χi = 1
places a constraint on the last mole fraction. As such, there are C – 1 compositional degrees of freedom for each phase present,
where C is the number of components in the mixture. Similarly, all but one of the chemical potentials of each phase present must
be equal, leaving only one that can be varied independently, leading to P – 1 thermodynamic constraints placed on each component.
Finally, there are two state variables that can be varied (such as pressure and temperature), adding two additional degrees of
freedom to the system. The net number of degrees of freedom is determined by adding all of the degrees of freedom and
subtracting the number of thermodynamic constraints.
F = 2 + P (C − 1) − C (P − 1)
= 2 +PC −P −PC +C
= 2 +C −P (8.3.1)
Example 8.3.1:
Show that the maximum number of phases that can co-exist at equilibrium for a single component system is P =3 .
Solution
The maximum number of components will occur when the number of degrees of freedom is zero.
0 = 2 +1 −P
P =3
Note: This shows that there can never be a “quadruple point” for a single component system!
Because a system at its triple point has no degrees of freedom, the triple point makes a very convenient physical condition at which
to define a temperature. For example, the International Practical Temperature Scale of 1990 (IPT-90) uses the triple points of
hydrogen, neon, oxygen, argon, mercury, and water to define several low temperatures. (The calibration of a platinum resistance
thermometer at the triple point of argon, for example, is described by Strouse (Strouse, 2008)). The advantage to using a triple
point is that the compound sets both the temperature and pressure, rather than forcing the researcher to set a pressure and then
measure the temperature of a phase change, introducing an extra parameter than can introduce uncertainty into the measurement.
This page titled 8.3: Criterion for Phase Equilibrium is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
8.3.1 https://chem.libretexts.org/@go/page/84338
8.4: The Clapeyron Equation
Based on the thermodynamic criterion for equilibrium, it is possible to draw some conclusions about the state variables p and T
and how they are related along phase boundaries. First, the chemical potentials of the two phases α and β in equilibrium with one
another must be equal.
μα = μβ (8.4.1)
Also, any infinitesimal changes to the chemical potential of one phase must be offset by an infinitesimal change to the chemical
potential of the other phase that is equal in magnitude.
Taking the difference between these Equations 8.4.1 and 8.4.2 shows that
dμα = dμβ
And since dμ can be expressed in terms of molar volume and molar entropy
dμ = V dp − SdT
It is clear that there will be constraints placed on changes of temperature and pressure while maintaining equilibrium between the
phases.
Vα dP − Sα dT = Vβ dP − Sβ dT
Gathering pressure terms on one side and temperature terms on the other
ΔV dp = ΔSdT
or
dp ΔS
= (8.4.3)
dT ΔV
Equation 8.4.3 is the Clapeyron equation. This expression makes it easy to see how the phase diagram for water is qualitatively
different than that for most substances. Specifically, the negative slope of the solid-liquid boundary on a pressure-temperature
phase diagram for water is very unusual, and arises due to the fact that for water, the molar volume of the liquid phase is smaller
than that of the solid phase.
Given that for a phase change
ΔHphase
ΔSphase =
T
= 0.9167 g/c m while that for liquid water is ρ = 0.9999 g/c m ) for an increase in pressure of 1.00 atm at 273 K.
3 3
ρice liquid
Solution
The molar volume of ice is given by
8.4.1 https://chem.libretexts.org/@go/page/84339
3
g 1 mol 1000 cm L
(0.9167 )( )( ) = 50.88
3
cm 18.016 g 1L mol
o
The molar volume of liquid water at 0 C is given by
3
g 1 mol 1000 cm L
(0.9999 )( )( ) = 55.50
3
cm 18.016 g 1L mol
So ΔV for the phase change of solid → liquid (which corresponds to an endothermic change) is
L L L
50.88 − 55.50 = −4.62
mol mol mol
To find the change in temperature, use the Clapeyron Equation (Equation 8.4.4) and separating the variables
ΔHf us dt
dp =
ΔV T
Integration (with the assumption that ΔH f us /ΔV does not change much over the temperature range) yields
p2 T2
ΔHf us dt
∫ dp = ∫
p1
ΔV T1
T
ΔHf us T2
p2 − p1 = Δp = ln( )
ΔV T1
or
ΔV Δp
T2 = T1 exp ( )
ΔHf us
so
L
⎛ ⎞
(1 atm) (−4.62 )
⎜ mol 8.314 J ⎟
T2 = (273 K) exp ⎜ ( )⎟
⎜ ⎟
⎜ J 0.08206 atm L ⎟
6009
⎝ mol ⎠
conversion factor
T2 = 252.5 K
So the melting point will decrease by 20.5 K. Note that the phase with the smaller molar volume is favored at the higher
pressure (as expected from Le Chatelier's principle)!
This page titled 8.4: The Clapeyron Equation is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
8.4.2 https://chem.libretexts.org/@go/page/84339
8.5: The Clausius-Clapeyron Equation
The Clapeyron equation can be developed further for phase equilibria involving the gas phase as one of the phases. This is the case
for either sublimation (solid → gas) or vaporization (liquid → gas). In the case of vaporization, the change in molar volume can
be expressed
ΔV = Vgas − Vliquid
Since substances undergo a very large increase in molar volume upon vaporization, the molar volume of the condensed phase
(liquid in this case) is negligibly small compared to the molar volume of the gas (i.e., V ≫ V ). So, gas liquid
ΔV ≈ Vgas
Noting that
dT 1
= −d ( )
2
T T
makes the integration very easy. If the enthalpy of vaporization is independent of temperature over the range of conditions,
p2 T2
dp ΔHvap 1
∫ =− ∫ d( )
p p R T1 T
1
p2 ΔHvap 1 1
ln( ) =− ( − ) (8.5.2)
p1 R T2 T1
This is the Clausius-Clapeyron equation. It can also be used to describe the boundary between solid and vapor phases by
substituting the enthalpy of sublimation (ΔH )sub
Example 8.5.1
The vapor pressure of a liquid triples when the temperature is increased from 25 °C to 45 °C. What is the enthalpy of
vaporization for the liquid?
Solution
The problem can be solved using the Clausius-Clapeyron equation (Equation 8.5.2). The following values can be used:
p2 = 3 p1 T2 = 318 K
p1 = p1 T1 = 298 K
8.5.1 https://chem.libretexts.org/@go/page/84340
3p1 ΔHvap 1 1
ln( ) =− ( − )
p1 J 318 K 298 K
9.314
mol K
J kJ
ΔHvap = 43280 = 43.28
mol mol
The Clausius-Clapeyron equation also suggests that a plot of ln(p) vs. 1/T should yield a straight line, the slope of which is
– ΔH /R (provided that ΔH is independent of temperature over the range of temperatures involved..
vap
ΔHvap 1
ln(p) = − ( ) + const.
R T
This approach in Example 8.5.1 is very useful when there are several pairs of measurements of vapor pressure and temperature.
Such a plot is shown below for water.
Figure 8.5.1
For water, which has a very large temperature dependence, the linear relationship of ln(p) vs. 1/T holds fairly well over a broad
range of temperatures. So even though there is some curvature to the data, a straight line fit still results in a reasonable description
of the data (depending, of course, on the precision needed in the experiment.) For this fit of the data, ΔH is found to be 43.14 vap
kJ/mol.
For systems that warrant it, temperature dependence of ΔH can be included into the derivation of the model to fit vapor
vap
pressure as a function of temperature. For example, if the enthalpy of vaporization is assumed to take the following empirical form
2
ΔHvap = ΔHo + aT + b T
and substituting it into the differential form of the Clausius-Clapeyron equation (Equation 8.5.1) generates
2
dp ΔHo + aT + b T dT
=
2
p R T
or
dp ΔHo dT a dT b
= + + dT
2
p R T R T R
The results of fitting these data to the temperature dependent model are shown in the table below.
8.5.2 https://chem.libretexts.org/@go/page/84340
ΔH0 (J mol-1) a (J mol-1 K-1) b (J mol-1 K-2) c
This results in calculated values of ΔH of 43.13 kJ/mol at 298 K, and 43.15 kJ/mol at 373 K. The results are a little bit skewed
vap
since there is no data above 100 oC included in the fit. A larger temperature dependence would be found if the higher-temperature
data were included in the fit.
This page titled 8.5: The Clausius-Clapeyron Equation is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
8.5.3 https://chem.libretexts.org/@go/page/84340
8.6: Phase Diagrams for Binary Mixtures
As suggested by the Gibbs Phase Rule, the most important variables describing a mixture are pressure, temperature and
composition. In the case of single component systems, composition is not important so only pressure and temperature are typically
depicted on a phase diagram. However, for mixtures with two components, the composition is of vital important, so there is
generally a choice that must be made as to whether the other variable to be depicted is temperature or pressure.
Temperature-composition diagrams are very useful in the description of binary systems, many of which will for two-phase
compositions at a variety of temperatures and compositions. In this section, we will consider several types of cases where the
composition of binary mixtures are conveniently depicted using these kind of phase diagrams.
Figure 8.6.1 :
As is the case for most solutes, their solubility is dependent on temperature. For many binary mixtures of immiscible liquids,
miscibility increases with increasing temperature. And then at some temperature (known as the upper critical temperature), the
liquids become miscible in all compositions. An example of a phase diagram that demonstrates this behavior is shown in Figure
8.6.1. An example of a binary combination that shows this kind of behavior is that of methyl acetate and carbon disufide, for which
the critical temperature is approximately 230 K at one atmosphere (Ferloni & Spinolo, 1974). Similar behavior is seen for
hexane/nitrobenzene mixtures, for which the critical temperature is 293 K.
Figure 8.6.2 :
8.6.1 https://chem.libretexts.org/@go/page/84341
Another condition that can occur is for the two immiscible liquids to become completely miscible below a certain temperature, or
to have a lower critical temperature. An example of a pair of compounds that show this behavior is water and trimethylamine. A
typical phase diagram for such a mixture is shown in Figure 8.6.2. Some combinations of substances show both an upper and lower
critical temperature, forming two-phase liquid systems at temperatures between these two temperatures. An example of a
combination of substances that demonstrate the behavior is nicotine and water.
Figure 8.6.3 :
Suppose that the temperature and composition of the mixture is given by point b in the above diagram. The horizontal line segment
that passes through point b, is terminated at points a and c, which indicate the compositions of the two liquid phases. Point a
indicates the mole faction of compound B (χ ) in the layer that is predominantly A, whereas the point c indicates the composition
A
B
(χ )of the layer that is predominantly compound B. The relative amounts of material in the two layers is then inversely
B
B
proportional to the length of the tie-lines a-b and b-c, which are given by l and l respectively. In terms of mole fractions,
A B
A
lA = χB − χ
B
and
B
lA = χ − χB
B
The number of moles of material in the A layer (n ) and the number of moles in the B layer (n ) are inversely proportional to the
A B
nA lA = nB lB
Or, substituting the above definitions of the lengths lA and lB , the ratio of these two lengths gives the ratio of moles in the two
phases.
B
nA lB χ − χB
B
= =
A
nB lA χB − χ
B
This page titled 8.6: Phase Diagrams for Binary Mixtures is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated
by Patrick Fleming.
8.6.2 https://chem.libretexts.org/@go/page/84341
8.7: Liquid-Vapor Systems - Raoult’s Law
Liquids tend to be volatile, and as such will enter the vapor phase when the temperature is increased to a high enough value
(provided they do not decompose first!) A volatile liquid is one that has an appreciable vapor pressure at the specified temperature.
An ideal mixture continuing at least one volatile liquid can be described using Raoult’s Law.
Raoult’s Law
Raoult’s law can be used to predict the total vapor pressure above a mixture of two volatile liquids. As it turns out, the composition
of the vapor will be different than that of the two liquids, with the more volatile compound having a larger mole fraction in the
vapor phase than in the liquid phase. This is summarized in the following theoretical diagram for an ideal mixture of two
compounds, one having a pure vapor pressure of p = 450 T orr and the other having a pure vapor pressure of p = 350 T orr . In
o
B
o
B
Figure 8.7.1, the liquid phase is represented at the top of the graph where the pressure is higher.
Figure 8.7.1 : The liquid phase is represented at the top of the graph where the pressure is higher
Oftentimes, it is desirable to depict the phase diagram at a single pressure so that temperature and composition are the variables
included in the graphical representation. In such a diagram, the vapor, which exists at higher temperatures) is indicated at the top of
the diagram, while the liquid is at the bottom. A typical temperature vs. composition diagram is depicted in Figure 8.7.2 for an
ideal mixture of two volatile liquids.
A B
o
by χ has its temperature increased to that indicated by point c, The system will consist of two phases, a liquid phase, with a
c
B
composition indicated by χ and a vapor phase indicated with a composition indicated by χ . The relative amounts of material in
d
B
b
B
the new vapor will have the composition consistent with χ . This demonstrates how the more volatile liquid (the one with the
a
B
lower boiling temperature, which is A in the case of the above diagram) can be purified from the mixture by collecting and re-
8.7.1 https://chem.libretexts.org/@go/page/84342
evaporating fractions of the vapor. If the liquid was the desired product, one would collect fractions of the residual liquid to achieve
the desired result. This process is known as distillation.
This page titled 8.7: Liquid-Vapor Systems - Raoult’s Law is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by Patrick Fleming.
8.7.2 https://chem.libretexts.org/@go/page/84342
8.8: Non-ideality - Henry's Law and Azeotropes
The proceeding discussion was based on the behaviors of ideal solutions of volatile compounds, and for which both compounds
follow Raoult’s Law. Henry’s Law can be used to describe these deviations.
o
pB = kH p
B
For which the Henry’s Law constant (k ) is determined for the specific compound. Henry’s Law is often used to describe the
H
solubilities of gases in liquids. The relationship to Raoult’s Law is summarized in Figure 8.8.1.
Figure 8.8.1 : The relationship between Raoult's Law and Henry's Law for a binary mixture.
Henry’s Law is depicted by the upper straight line and Raoult’s Law by the lower.
The solubility of C O (g) in water at 25 oC is 3.32 x 10-2 M with a partial pressure of C O over the solution of 1 bar.
2 2
Assuming the density of a saturated solution to be 1 kg/L, calculate the Henry’s Law constant for C O . 2
Solution
In one L of solution, there is 1000 g of water (assuming the mass of CO2 dissolved is negligible.)
1 mol
(1000 g) ( ) = 55 mol H2 O
18.02 g
The solubility of C O can be used to find the number of moles of C O dissolved in 1 L of solution also:
2 2
−2
3.32 × 10 mol
−2
⋅ 1 L = 3.32 × 10 mol C O2
L
−2
3.32 × 10 mol
−4
χb = = 5.98 × 10
55.5 mol
And so
5 −4
10 P a = 5.98 × 10 kH
or
9
kH = 1.67 × 10 Pa
Azeotropes
An azeotrope is defined as the common composition of vapor and liquid when they have the same composition.
8.8.1 https://chem.libretexts.org/@go/page/84538
Figure 8.8.2 : Phase diagrams for (left) a maximum boiling point azeotrope and (right) Ta maximum boiling point azeotrope.
Azeotropes can be either maximum boiling or minimum boiling, as show in Figure 8.8.2; lef t. Regardless, distillation cannot
purify past the azeotrope point, since the vapor and quid phases have the same composition. If a system forms a minimum boiling
azeotrope and also has a range of compositions and temperatures at which two liquid phases exist, the phase diagram might look
like Figure 8.8.2; right:
Figure 8.8.3 : Phase diagram for a binary solution with the boiling point of a minimum boiling azeotrope that is higher that when
components are miscible (single phase).
Another possibility that is common is for two substances to form a two-phase liquid, form a minimum boiling azeotrope, but for the
azeotrope to boil at a temperature below which the two liquid phases become miscible. In this case, the phase diagram will look
like Figure 8.8.3.
8.8.2 https://chem.libretexts.org/@go/page/84538
Example 8.8.1:
In the diagram, make up of a system in each region is summarized below the diagram. The point e indicates the azeotrope
composition and boiling temperature.
Solution
Within each two-phase region (III, IV, and the two-phase liquid region, the lever rule will apply to describe the composition of
each phase present. So, for example, the system with the composition and temperature represented by point b (a single-phase
liquid which is mostly compound A, designated by the composition at point a, and vapor with a composition designated by that
at point c), will be described by the lever rule using the lengths of tie lines lA and lB.
This page titled 8.8: Non-ideality - Henry's Law and Azeotropes is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by Patrick Fleming.
8.8.3 https://chem.libretexts.org/@go/page/84538
8.9: Solid-Liquid Systems - Eutectic Points
A phase diagram for two immiscible solids and the liquid phase (which is miscible in all proportions) is shown in Figure 8.9.1. The
point labeled “e2” is the eutectic point, meaning the composition for which the mixture of the two solids has the lowest melting
point. The four main regions can be described as below:
I. Two-phase solid
II. Solid (mostly A) and liquid (A and B)
III. Solid (mostly B) and liquid (A and B)
IV. Single phase liquid (A and B)
Figure 8.9.1 : Phase diagram of a two-component system that exhibits an eutectic point.
The unlabeled regions on the sides of the diagram indicate regions where one solid is so miscible in the other, that only a single
phase solid forms. This is different than the “two-phase solid” region where there are two distinct phases, meaning there are regions
(crystals perhaps) that are distinctly A or B, even though they are intermixed within on another. Region I contains two phases: a
solid phase that is mostly compound A, and a liquid phase which contains both A and B. A sample in region II (such as the
temperature/composition combination depicted by point b) will consist of two phases: 1 is a liquid mixture of A and B with a
composition given by that at point a, and the other is a single phase solid that is mostly pure compound B, but with traces of A
entrained within it. As always, the lever rule applies in determining the relative amounts of material in the two phases.
In the case where the widths of the small regions on either side of the phase diagram are negligibly small, a simplified diagram with
a form similar to that shown in Figure 8.9.2 can be used. In this case, it is assumed that the solids never form a single phase! The
tin-lead system exhibits such behavor.
Figure 8.9.2 : A simplified phase diagram of a two-component system that exhibits an eutectic point.
Another important case is that for which the two compounds A and B can react to form a third chemical compound C. If the
compound C is stable in the liquid phase (does not decompose upon melting), the phase diagram will look like Figure 8.9.3.
8.9.1 https://chem.libretexts.org/@go/page/84542
Figure 8.9.3 : A simplified phase diagram of a two-component system that exhibits an eutectic point.
In this diagram, the vertical boundary at χ = 0.33 is indicative of the compound C formed by A and B . From the mole fraction
B
of B , it is evident that the formula of compound C is A B . The reaction that forms compound C is
2
2A + B → C
Thus, at overall compositions where χ < 0.33, there is excess compound A (B is the limiting reagent) and for χ there is an
B B
excess of compound B (A is now the limiting reagent.) With this in mind, the makeup of the sample in each region can be
summarized as
I. Two phase solid (A and C)
II. Two phase solid (C and B)
III. Solid A and liquid (A and C)
IV. Solid C and liquid (A and C)
V. Solid C and liquid (C and B)
VI. Solid B and liquid (C and B)
VII. liquid. Single phase liquid (A and C or C and B, depending on which is present in excess)
Zinc and Magnesium are an example of two compounds that demonstrate this kind of behavior, with the third compound having the
formula Z n M g (Ghosh, Mezbahul-Islam, & Medraj, 2011).
2
Incongruent Melting
Oftentimes, the stable compound formed by two solids is only stable in the solid phase. In other words, it will decompose upon
melting. As a result, the phase diagram will take a lightly different form, as is shown in Figure 8.9.4.
Figure 8.9.4 : A phase diagram of a two-component system that exhibits incongruent melting.
In this diagram, the formula of the stable compound is AB (consistent with χ < 0.75). But you will notice that the boundary
3 B
separating the two two-phase solid regions does not extend all of the way to the single phase liquid portion of the diagram. This is
8.9.2 https://chem.libretexts.org/@go/page/84542
because the compound will decompose upon melting. The process of decomposition upon melting is also called incongruent
melting. The makeup of each region can be summarized as
I. Two phase solid (A and C)
II. Two phase solid (C and B)
III. Solid A and liquid (A and B)
IV. Solid C and liquid (A and B)
V. Solid B and liquid (A and B)
There are many examples of pairs of compounds that show this kind of behavior. One combination is sodium and potassium, which
form a compound (N a K ) that is unstable in the liquid phase and so it melts incongruently (Rossen & Bleiswijk, 1912).
2
This page titled 8.9: Solid-Liquid Systems - Eutectic Points is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by Patrick Fleming.
8.9.3 https://chem.libretexts.org/@go/page/84542
8.10: Cooling Curves
The method that is used to map the phase boundaries on a phase diagram is to measure the rate of cooling for a sample of known
composition. The rate of cooling will change as the sample (or some portion of it) begins to undergo a phase change. These
“breaks” will appear as changes in slope in the temperature-time curve. Consider a binary mixture for which the phase diagram is
as shown in Figure 8.10.1A. A cooling curve for a sample that begins at the temperature and composition given by point a is
shown in Figure 8.10.1B.
Figure 8.10.1 : (A) cooling of a two-component system from liquid to solid. (B) Cooresponding cooling curve for this process.
As the sample cools from point a, the temperature will decrease at a rate determined by the sample composition, and the geometry
of the experiment (for example, one expects more rapid cooling is the sample has more surface area exposed to the cooler
surroundings) and the temperature difference between the sample and the surroundings.
When the temperature reaches that at point b, some solid compound B will begin to form. This will lead to a slowing of the cooling
due to the exothermic nature of solid formation. But also, the composition of the liquid will change, becoming richer in compound
A as B is removed from the liquid phase in the form of a solid. This will continue until the liquid attains the composition at the
eutectic point (point c in the diagram.)
When the temperature reaches that at point c, both compounds A and B will solidify, and the composition of the liquid phase will
remain constant. As such, the temperature will stop changing, creating what is called the eutectic halt. Once all of the material has
solidified (at the time indicated by point c’), the cooling will continue at a rate determined by the heat capacities of the two solids A
and B, the composition, and (of course) the geometry of the experimental set up. By measuring cooling curves for samples of
varying composition, one can map the entire phase diagram.
This page titled 8.10: Cooling Curves is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick Fleming.
8.10.1 https://chem.libretexts.org/@go/page/84543
Welcome to the Chemistry Library. This Living Library is a principal hub of the LibreTexts project, which is a multi-institutional
collaborative venture to develop the next generation of open-access texts to improve postsecondary education at all levels of higher
learning. The LibreTexts approach is highly collaborative where an Open Access textbook environment is under constant revision
by students, faculty, and outside experts to supplant conventional paper-based books.
1 https://chem.libretexts.org/@go/page/84541
8.S: Phase Equilibrium (Summary)
Learning Objectives
After mastering the material in this chapter, one will be able to
1. State the thermodynamic criterion for equilibrium in terms of chemical potential.
2. Derive and interpret the Gibbs Phase Rule.
3. Derive the Clapeyron equation from the thermodynamic criterion for equilibrium.
4. Interpret the slope of phase boundaries on a pressure-temperature phase diagram in terms of the relevant changes in entropy and
molar volume for the given phase change.
5. Derive the Clausius-Clapeyron equation, stating all of the necessary approximations.
6. Use the Clausius-Clapeyron equation to calculate the vapor pressure of a substance or the enthalpy of a phase change from
pressure-temperature data.
7. Interpret phase diagrams for binary mixtures, identifying the phases and components present in each region.
8. Perform calculations using Raoult’s Law and Henry’s Law to relate vapor pressure to composition in the liquid phase.
9. Describe the distillation process, explaining how the composition of liquid and vapor phases can differ, and how azeotrope
composition place bottlenecks in the distillation process.
10. Describe how cooling curves are used to derive phase diagrams by locating phase boundaries.
References
1. Ferloni, P., & Spinolo, G. (1974). Int. DATA Ser., Sel. Data Mixtures, Ser. A, 70.
2. Ghosh, P., Mezbahul-Islam, M., & Medraj, M. (2011). Critical assessment and thermodynamic modeling of Mg–Zn, Mg–Sn,
Sn–Zn and Mg–Sn–Zn systems. Calphad, 36, 28-43.
3. Nave, R. (n.d.). Saturated Vapor Pressure, Density for Water. (Georgia State University, Department of Physics and Astronomy)
Retrieved April 7, 2016, from HyperPhysics: http://hyperphysics.phy-astr.gsu.edu...ic/watvap.html
4. Rossen, G. L., & Bleiswijk, H. v. (1912). Über das Zustandsdiagramm der Kalium-Natriumlegierungen. Zeitschrift für
anorganische Chemie, 74(1), 152-156.
5. Strouse, G. F. (2008, January). Standard Platinum Resistance Thermometer Calibrations from the Ar TP to the Ag FP. National
Institute of Standards and Technology Special Publication 250-81.
8.S.1 https://chem.libretexts.org/@go/page/84540
This page titled 8.S: Phase Equilibrium (Summary) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
8.S.2 https://chem.libretexts.org/@go/page/84540
CHAPTER OVERVIEW
9: Chemical Equilibria
This page titled 9: Chemical Equilibria is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
1
9.1: Prelude to Chemical Equilibria
The small is great, the great is small; all is in equilibrium in necessity... - Victor Hugo in
“Les Miserables”
As was discussed in Chapter 6, the natural tendency of chemical systems is to seek a state of minimum Gibbs function. Once the
minimum is achieved, movement in any chemical direction will not be spontaneous. It is at this point that the system achieves a
state of equilibrium.
From the diagram above, it should be clear that the direction of spontaneous change is determined by minimizing
∂G
( ) .
∂ξ
p,T
If the slope of the curve is negative, the reaction will favor a shift toward products. And if it is positive, the reaction will favor a
shift toward reactants. This is a non-trivial point, as it underscores the importance of the composition of the reaction mixture in the
determination of the direction of the reaction.
This page titled 9.1: Prelude to Chemical Equilibria is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
9.1.1 https://chem.libretexts.org/@go/page/84344
9.2: Chemical Potential
Equilibrium can be understood as accruing at the composition of a reaction mixture at which the aggregate chemical potential of
the products is equal to that of the reactants. Consider the simple reaction
A(g) ⇌ B(g)
μA = μB
If the gases behave ideally, the chemical potentials can be described in terms of the mole fractions of A and B
pA pB
o o
μ + RT ln( ) =μ + RT ln( ) (9.2.1)
A B
ptot ptot
where Dalton’s Law has been used to express the mole fractions.
pi
χi =
ptot
Equation 9.2.1 can be simplified by collecting all chemical potentials terms on the left
pB pA
o o
μ −μ = RT ln( ) − RT ln( ) (9.2.2)
A B
ptot ptot
And since p A / pB = Kp for this reaction (assuming perfectly ideal behavior), one can write
o
ΔG = RT ln Kp
Another way to achieve this result is to consider the Gibbs function change for a reaction mixture in terms of the reaction
quotient. The reaction quotient can be expressed as
νi
∏ p
i i
Qp =
νj
∏ p
j j
where ν are the stoichiometric coefficients for the products, and ν are those for the reactants. Or if the stoichiometric coefficients
i j
0 = ∑ νi Xi
where X refers to one of the species in the reaction, and ν is then the stoichiometric coefficient for that species, it is clear that ν
i i i
will be negative for a reactant (since its concentration or partial pressure will reduce as the reaction moves forward) and positive
for a product (since the concentration or partial pressure will be increasing.) If the stoichiometric coefficients are expressed in this
way, the expression for the reaction quotient becomes
νi
Qp = ∏ p
i
Using this expression, the Gibbs function change for the system can be calculated from
o
ΔG = ΔG + RT ln Qp
9.2.1 https://chem.libretexts.org/@go/page/84345
ΔG = 0
and
Qp = Kp
It is evident that
o
ΔGrxn = −RT ln Kp (9.2.3)
It is also of value to note that the criterion for a spontaneous chemical process is that ΔG < 0 , rather than ΔG , as is
rxn
o
rxn
fugacity or activity. However, the direction of spontaneous change for a chemical reaction is dependent on the composition of
the reaction mixture. Similarly, the magnitude of the equilibrium constant is insufficient to determine whether a reaction will
spontaneously form reactants or products, as the direction the reaction will shift is also a function of not just the equilibrium
constant, but also the composition of the reaction mixture!
Example 9.2.1:
Based on the data below at 298 K, calculate the value of the equilibrium constant (K ) for the reaction
p
N O(g) N O2 (g)
G
o
f
(kJ/mol) 86.55 51.53
Solution
First calculate the value of ΔG o
rxn from the ΔG data.
o
f
o
ΔGrxn = 2 × (51.53 kJ/mol) − 2 × (86.55 kJ/mol) = −70.04 kJ/mol
12
Kp = 1.89 × 10
This page titled 9.2: Chemical Potential is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
9.2.2 https://chem.libretexts.org/@go/page/84345
9.3: Activities and Fugacities
To this point, we have mostly ignored deviations from ideal behavior. But it should be noted that thermodynamic equilibrium
constants are not expressed in terms of concentrations or pressures, but rather in terms of activities and fugacities (both being
discussed in Chapter 7). Based on these quantities,
νi
Kp = ∏ f (9.3.1)
i
and
νi
Kc = ∏ a
i
And since activities and fugacities are unitless, thermodynamic equilibrium constants are unitless as well. Further, it can be noted
that the activities of solids and pure liquids are unity (assuming ideal behavior) since they are in their standard states at the given
temperature. As such, these species never change the magnitude of the equilibrium constant and are generally omitted from the
equilibrium constant expression.
Kp and Kc
Oftentimes it is desirable to express the equilibrium constant in terms of concentrations (or activities for systems that deviate from
ideal behavior.) To make this conversion, the relationship between pressure and concentration from the ideal gas law can be used.
n
p = RT ( )
V
And noting that the concentration is given by (n/V ), the expression for the equilibrium constant (Equation 9.3.1) becomes
νi
Kp = ∏(RT [ Xi ] ) (9.3.2)
And since for a given temperature, RT is a constant and can be factored out of the expression, leaving
νi νi
Kp = ( ∏(RT ) ) ( ∏[ Xi ] ) (9.3.3)
i i
∑ νi νi
= (RT ) ∏[ Xi ] (9.3.4)
∑ νi
= (RT ) Kc (9.3.5)
This conversion works for reactions in which all reactants and products are in the gas phase. Care must be used when applying this
relationship to heterogeneous equilibria.
This page titled 9.3: Activities and Fugacities is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
9.3.1 https://chem.libretexts.org/@go/page/84346
9.4: Pressure Dependence of Kp - Le Châtelier's Principle
Since the equilibrium constant K is a function of ΔG
p which is defined for a specific composition (all reactants in their
o
rxn
standard states and at unit pressure (or fugacity), changes in pressure have no effect on equilibrium constants for a fixed
temperature. However, changes in pressure can have profound effects on the compositions of equilibrium mixtures.
To demonstrate the relationship, one must recall Dalton’s law of partial pressures. According to this relationship, the partial
pressure of a component of a gas-phase mixture can be expressed
pi = χt ptot
It is the combination of mole fractions that describes the composition of the equilibrium mixture.
Substituting the above expression into the expression for K yields
p
νi
Kp = ∏(χi ptot )
This expression can be factored into two pieces – one containing the mole fractions and thus describing the composition, and one
containing the total pressure.
νi νi
Kp = ( ∏ χ ) (∏ p )
i tot
i i
The second factor is a constant for a given total pressure. If the first term is given the symbol K , the expression becomes
x
∑ νi
Kp = Kx (ptot ) i
∑ νi
i
Kx = ∏ χ
i
but is not itself a constant. The value ofK will vary with varying composition, and will need to vary with varying total pressure
x
Example 9.4.1:
Consider the following reaction at equilibrium.
In which direction will the equilibrium shift if the volume of the reaction vessel is decreased?
Solution
A decrease in the volume will lead to an increase in total pressure. Since the equilibrium constant can be expressed as
pc pD χp χD
−1
Kp = = (ptot )
2 2
pA p χA χ
B B
An increase in pressure will lead to an increase in Kx to maintain a constant value of Kp . So the reaction will shift to form
more of the products C and D.
Note: This should make some sense, since a shift to the side of the reaction with fewer moles of gas will lower the total
pressure of the reaction mixture, and thus relieving the stress introduced by increasing the pressure. This is exactly what is
expected according to Le Chatelier's principle.
It should be noted that there are several ways one can affect the total pressure of a gas-phase equilibrium. These include the
introduction or removal of reactants or products (perhaps through condensation or some other physical process), a change in
volume of the reaction vessel, or the introduction of an inert gas that does not participate in the reaction itself. (Changes in the
9.4.1 https://chem.libretexts.org/@go/page/84347
temperature will be discussed in a later section.) The principle of Le Chatelier's can be used as a guide to predict how the
equilibrium composition will respond to a change in pressure.
Example 9.4.2:
A 1.0 L vessel is charged with 1.00 atm of A, and the following reaction is allowed to come to equilibrium at 298 K.
A(g) ⇌ 2B(g)
with K p = 3.10 .
a. What are the equilibrium partial pressures and mole fractions of A and B?
b. If the volume of the container is doubled, what are the equilibrium partial pressures and mole fractions of A and B?
c. If 1.000 atm of Ar (an inert gas) is introduced into the system described in b), what are the equilibrium partial pressures and
mole fractions of A and B once equilibrium is reestablished?
Solution
Part a:
First, we can use an ICE[1] table to solve part a).
A 2B
Change -x +2x
2
(2x)
3.10 atm =
1.00 atm − x
x1 = −1.349 atm
x1 = 0.574 atm
Clearly, x , while a solution to the mathematical problem, is not physically meaningful since the equilibrium pressure of B
1
9.4.2 https://chem.libretexts.org/@go/page/84347
χB = 1 − χA = 1 − 0.271 = 0.729
Part b:
The volume is doubled. Again, an ICE table is useful. The initial pressures will be half of the equilibrium pressures found in
part a).
A 2B
Change -x +2x
x1 = −1.4077 atm
x1 = 0.05875 atm
We reject the negative root (since it would cause both of the partial pressures to become negative. So the new equilibrium
partial pressures are
pA = 0.154 atm
pB = 0.0692 atm
χA = 0.182
χB = 0.818
We can see that the mole fraction of A decreased and the mole fraction B increased. This is the result expected by Le
Chatlier’s principle since the lower total pressure favors the side of the reaction with more moles of gas.
Part c:
We introduce 1.000 atm of an inert gas. The new partial pressures are
pA = 0.154 atm
pB = 0.692 atm
And because the partial pressures of A and B are unaffected, the equilibrium does not shift! What is affected is the
composition, and so the mole fractions will change.
0.154 atm
χA = = 0.08341
0.154atm + 0.692 atm + 1.000 atm
0.692 atm
χB = = 0.08341
0.154atm + 0.692 atm + 1.000 atm
1.000 atm
χAr = = 0.08341
0.154atm + 0.692 atm + 1.000 atm
And since
Kp = Kx (ptot )
9.4.3 https://chem.libretexts.org/@go/page/84347
2
(0.3749)
(1.846 atm) = 3.1
0.08342
Within round-off error, the value obtained is the equilibrium constant. So the conclusion is that the introduction of an inert gas,
even though it increases the total pressure, does not induce a change in the partial pressures of the reactants and products, so it
does not cause the equilibrium to shift.
[1] ICE is an acronym for “Initial, Change, Equilibrium”. An ICE table is a tool that is used to solve equilibrium problems in terms
of an unknown number of moles (or something proportional to moles, such as pressure or concentration) will shift for a system to
establish equilibrium. See (Tro, 2014) or a similar General Chemistry text for more background and information.
This page titled 9.4: Pressure Dependence of Kp - Le Châtelier's Principle is shared under a CC BY-NC-SA 4.0 license and was authored,
remixed, and/or curated by Patrick Fleming.
9.4.4 https://chem.libretexts.org/@go/page/84347
9.5: Degree of Dissociation
Reactions such as the one in the previous example involve the dissociation of a molecule. Such reactions can be easily described in
terms of the fraction of reactant molecules that actually dissociate to achieve equilibrium in a sample. This fraction is called the
degree of dissociation. For the reaction in the previous example
A(g) ⇌ 2B(g)
the degree of dissociation can be used to fill out an ICE table. If the reaction is started with n moles of A , and a is the fraction of A
molecules that dissociate, the ICE table will look as follows.
A 2B
Initial n 0
1 −α
=
1 +α
2α
χB =
1 +α
1 +α
2
4α
= (9.5.2)
2
1 −α
∑ νi
Kp = Kx (ptot ) (9.5.3)
is given by
2
4α
Kp = (ptot )
2
(1 − α )
Example 9.5.1
Based on the values given below, find the equilibrium constant at 25 oC and degree of dissociation for a system that is at a total
pressure of 1.00 atm for the reaction
N2 O4 (g) ⇌ 2N O2 (g)
N2 O4 (g) N O2 (g)
ΔG
o
f
(kJ/mol) 99.8 51.3
Solution
9.5.1 https://chem.libretexts.org/@go/page/84348
First, the value of K can be determined from ΔG
p
o
rxn via an application of Hess' Law.
o
ΔGrxn = 2 (51.3 kJ/mol) − 99.8 kJ/mol = 2.8 kJ/mol
Kp = 0.323 atm
The degree of dissociation can then be calculated from the ICE tables at the top of the page for the dissociation of N 2 O4 (g) :
2
4α
Kp = (ptot )
1 − α2
2
4α
0.323 atm = (1.00 atm)
2
1 −α
Solving for α ,
α = 0.273
Note: since a represents the fraction of N2O4 molecules dissociated, it must be a positive number between 0 and 1.
Example 9.5.2
A + 2B ⇌ 2C
A reaction vessel is initially filled with 1.00 mol of A and 2.00 mol of B. At equilibrium, the vessel contains 0.60 mol C and a
total pressure of 0.890 atm at 1350 K.
1. How many mol of A and B are present at equilibrium?
2. What is the mole fraction of A, B, and C at equilibrium?
3. Find values for K , K , and ΔG .
x p
o
rxn
Solution
Let’s build an ICE table!
A 2B 2C
From the equilibrium measurement of the number of moles of C, x = 0.30 mol. So at equilibrium,
A 2B 2C
The total number of moles at equilibrium is 2.70 mol. From these data, the mole fractions can be determined.
9.5.2 https://chem.libretexts.org/@go/page/84348
0.70 mol
χA = = 0.259
2.70 mol
1.40 mol
χB = = 0.519
2.70 mol
0.60 mol
χC = = 0.222
2.70 mol
So K is given by
x
2
(0.222)
Kx = = 0.7064
2
(0.259)(0.519)
−1 −1
Kp = 0.7604(0.890 atm ) = 0.792 atm
The thermodynamic equilibrium constant is unitless, of course, since the pressures are all divided by 1 atm. So the actual value of
K is 0.794. This value can be used to calculate ΔG using
o
p rxn
o
ΔGrxn = −RT ln Kp
so
o
ΔGrxn = −(8.314 J/(mol K))(1350 K) ln(0.792)
= 2590 J/mol
This page titled 9.5: Degree of Dissociation is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
9.5.3 https://chem.libretexts.org/@go/page/84348
9.6: Temperature Dependence of Equilibrium Constants - the van ’t Hoff Equation
The value of K is independent of pressure, although the composition of a system at equilibrium may be very much dependent on
p
pressure. Temperature dependence is another matter. Because the value of ΔG is dependent on temperature, the value of K is
o
rxm p
as well. The form of the temperature dependence can be taken from the definition of the Gibbs function. At constant temperature
and pressure
o o
ΔG ΔG 1 1
T2 T1 o
− = ΔH ( − )
T2 T1 T2 T1
Substituting
o
ΔG = −RT ln K
For the two values of ΔG and using the appropriate temperatures, yields
o
−RT2 ln K2 −RT1 ln K1 o
1 1
− = ΔH ( − )
T2 T1 T2 T1
And simplifying the expression so that only terms involving K are on the left and all other terms are on the right results in the van
’t Hoff equation, which describes the temperature dependence of the equilibrium constant.
o
K2 ΔH 1 1
ln( ) =− ( − ) (9.6.1)
K1 R T2 T1
Because of the assumptions made in the derivation of the Gibbs-Helmholtz equation, this relationship only holds if ΔH is o
independent of temperature over the range being considered. This expression also suggests that a plot of ln(K) as a function of
1/T should produce a straight line with a slope equal to – ΔH /R. Such a plot is known as a van ’t Hoff plot, and can be used to
o
Example 9.6.1
Solution
This is a job for the van ’t Hoff equation!
T1 = 298 K
T2 = 310 K
o
ΔHrxm = 32.4 kJ/mol
K1 = 0.0260
K2 = ?
So Equation 9.6.1 becomes
K2 32400 J/mol 1 1
ln( ) =− ( − )
0.0260 8.314 K/(mol K) 310 K 298 K
K2 = 0.0431
Note: the value of K increased with increasing temperature, which is what is expected for an endothermic reaction. An
2
increase in temperature should result in an increase of product formation in the equilibrium mixture. But unlike a change in
pressure, a change in temperature actually leads to a change in the value of the equilibrium constant!
9.6.1 https://chem.libretexts.org/@go/page/84349
Example 9.6.2
Given the following average bond enthalpies for P−Cl and Cl−Cl bonds, predict whether or not an increase in temperature
will lead to a larger or smaller degree of dissociation for the reaction
PCl ⇌ PCl + Cl
5 3 2
P-Cl 326
Cl-Cl 240
Solution
The estimated reaction enthalpy is given by the total energy expended breaking bonds minus the energy recovered by the
formation of bonds. Since this reaction involves breaking two P-Cl bonds (costing 652 kJ/mol) and the formation of one Cl-Cl
bond (recovering 240 kJ/mol), it is clear that the reaction is endothermic (by approximately 412 kJ/mol). As such, an increase
in temperature should increase the value of the equilibrium constant, causing the degree of dissociation to be increased at the
higher temperature.
This page titled 9.6: Temperature Dependence of Equilibrium Constants - the van ’t Hoff Equation is shared under a CC BY-NC-SA 4.0 license
and was authored, remixed, and/or curated by Patrick Fleming.
9.6.2 https://chem.libretexts.org/@go/page/84349
9.7: The Dumas Bulb Method for Measuring Decomposition Equilibrium
A classic example of an experiment that is employed in many physical chemistry laboratory courses uses a Dumas Bulb method to
measure the dissociation of N2O4(g) as a function of temperature (Mack & France, 1934). In this experiment, a glass bulb is used to
create a constant volume container in which a volatile substance can evaporate, or achieve equilibrium with other gases present.
The latter is of interest in the case of the reaction
N2 O4 (g) ⇌ 2N O2 (g) (9.7.1)
Figure 9.7.1 :
The Dumas bulb is then charged with a pure sample of the gas to be investigated (such as N2O4) and placed in a thermalized bath.
It is then allowed to come to equilibrium. Once Equilibrium is established, the stopcock is opened to allow gas to escape until the
internal pressure is set to the pressure of the room. The stopcock is then closed and the bulb weighed to determine the total mass of
gas remaining inside. The experiment is repeated at higher and higher temperatures (so that at each subsequent measurement, the
larger degree of dissociation creates more molecules of gas and an increase in pressure in the bulb (along with the higher
temperature), which then leads to the expulsion of gas when the pressure is equilibrated.
The degree of dissociation is then determined based on the calculated gas density at each temperature.
ρ1 − ρ2
α =
ρ2 (n − 1)
where ρ is the measured density and ρ is the theoretical density if no dissociation occurs (calculated from the ideal gas law for
1 2
the given temperature, pressure, and molar mass of the dissociating gas) and n is the number of fragments into which the
dissociating gas dissociates (ie.g., n = 2 for Equation \req{eq1}). The equilibrium constant is then calculated as
2
4α p
K = ( )
1 − α2 1.00 atm
This page titled 9.7: The Dumas Bulb Method for Measuring Decomposition Equilibrium is shared under a CC BY-NC-SA 4.0 license and was
authored, remixed, and/or curated by Patrick Fleming.
9.7.1 https://chem.libretexts.org/@go/page/84555
9.8: Acid-Base Equilibria
A great many processes involve proton transfer, or acid-base types of reactions. As many biological systems depend on carefully
controlled pH, these types of processes are extremely important. The pH is defined by
+ +
pH ≡ − log a(H ) ≈ − log[ H ] (9.8.1)
+ −
[H ][ A ]
Ka = (9.8.2)
[H A]
As is the case for all thermodynamic equilibrium constants, the concentrations are replaced by activities and the equilibrium
constant is unitless. However, if all species behave ideally (have unit activity coefficients) the units can be used as a very useful
guide in solving problems.
Solution
An ICE table will come in very handy here!
+ −
[HOAc] [H ] [OAc ]
Initial 0.200 M 0 0
Change -x +x +x
Equilibrium 0.200 M - x x x
This produces a quadratic equation, and thus two values of x which satisfy the relationship.
x1 = −0.001906 M
x2 = 0.001888 M
9.8.1 https://chem.libretexts.org/@go/page/84556
The Auto-ionization of Water
Water is a very important solvent as water molecules have large dipole moments which create favorable interactions with ionic
compounds. Water also has a large dielectric constant which damps the electric field generated by ions in solutions, making the
comparative interactions with water more favorable than with other ions in solution in many cases. But water also dissociates into
ions through the reaction
+ −
H2 O(l) ⇌ H (aq) + OH (aq) (9.8.3)
The equilibrium constant governing this dissociation is highly temperature dependent. The data below are presented by Bandura
and Lvov (Bandura & Lvov, 2006)
T (°C) 0 25 50 75 100
However, from the fit of these data, a value of ΔH can be determined to be 52.7 kJ/mol. Of particular note is that the
rxn
What is the pH of neutral water at 37 °C (normal human body temperature)? Neutral water no excess of [H +
] over [OH −
] or
vice versa.
Solution
From the best-fit line in the van’t Hoff plot of Figure 9.8.1, the value of K can be calculated:
w
6338 K
ln(Kw ) = − − 11.04
310 K
−14 2
Kw = 2.12 × 10 M
9.8.2 https://chem.libretexts.org/@go/page/84556
Note: This is slightly less than a pH of 7.00, which is normally considered to be “neutral.” But a pH of 7.00 is only neutral at
25 °C! At higher temperatures, neutral pH is a lower value due to the endothermic nature of the auto-ionization water. While it
has a nigher [H ] concentration, it also has a higher [OH ] and at the same level, so it is still technically neutral.
+ −
−
[H A][OH ]
Kb =
−
[A ]
The concentration (or activity) of the pure compound H2O is not included in the equilibrium expression because, being a pure
compound in its standard state, it has unit activity throughout the process of establishing equilibrium. Further, it should be noted
that when Kb is combined with the expression for Ka for the weak acid HA (Equation 9.8.2),
+ − −
[H ][ A ] [H A][OH ]
+ −
Ka Kb = ( )( ) = [H ][OH ] = Kw
−
[H A] [A ]
As a consequence, if one knows K for a weak acid, one also knows K for its conjugate base, since the product results in K .
a b w
Example 9.8.3:
What is the pH of a 0.150 M solution of KF? (For HF, pKa = 3.17 at 25 °C)
Solution
The problem involves the hydrolysis of the conjugate base of HF, F-. The hydrolysis reaction is
− −
F + H2 O ⇌ H F + OH
Initial 0.150 M 0 0
Change -x +x +x
Equilibrium 0.150 M - x x x
−14 2 2
Kw 1.0 × 10 M x
Kb = = =
3.17
Ka 10 M 1.50 M − x
In this case, the small value of K insures that the value of x will be negligibly small compared to 0.150 M. In this limit, the
b
So [H +
] is given by
−14 2
Kw 10 M
+ −9
[H ] = = = 6.71 × 10 M
− −6
[OH ] 1.49 × 10 M
9.8.3 https://chem.libretexts.org/@go/page/84556
−9
pH = − log (6.71 × 10 ) = 8.17
10
Note: The pH of this salt solution is slightly basic. This is to be expected as KF can be thought of being formed in the reaction
of a weak acid (HF) with a strong base (KOH). In the competition to control the pH, the strong base ends up winning the battle.
This page titled 9.8: Acid-Base Equilibria is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
9.8.4 https://chem.libretexts.org/@go/page/84556
9.9: Buffers
Buffer solutions, which are of enormous importance in controlling pH in various processes, can be understood in terms of acid/base
equilibrium. A buffer is created in a solution which contains both a weak acid and its conjugate base. This creates to absorb excess
H+ or supply H+ to replace what is lost due to neutralization. The calculation of the pH of a buffer is straightforward using an ICE
table approach.
Example 9.9.1:
What is the pH of a solution that is 0.150 M in KF and 0.250 M in HF?
Solution
The reaction of interest is
+ −
HF ⇌ H +F
Change -x +x +x
+ −
[H ][ F ]
Ka =
[H F ]
x(0.150 M + x)
−3.17
10 M =
0.250 M − x
This expression results in a quadratic relationship, leading to two values of x that will make it true. Rejecting the negative root, the
remaining root of the equation indicates
+
[H ] = 0.00111 M
So the pH is given by
For buffers made from acids with sufficiently large values of pKa the buffer problem can be simplified since the concentration of
the acid and its conjugate base will be determined by their pre-equilibrium values. In these cases, the pH can be calculated using
the Henderson-Hasselbalch approximation.
If one considers the expression for K a
+ − −
[H ][ A ] [H ]
+
Ka = = [H ]
[H A] [H A]
9.9.1 https://chem.libretexts.org/@go/page/84557
It should be noted that this approximation will fail if:
1. the pk is too small,
a
2. the concentrations [A −
] is too small, or
3. [H A] is too small,
since the equilibrium concentration will deviate wildly from the pre-equilibrium values under these conditions.
This page titled 9.9: Buffers is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick Fleming.
9.9.2 https://chem.libretexts.org/@go/page/84557
9.10: Solubility of Ionic Compounds
The solubility of ionic compounds in water can also be described using the concepts of equilibrium. If you consider the dissociation
of a generic salt MX
+ −
M X(s) ⇌ M (aq) + X (aq)
Ksp is the solubility product and is the equilibrium constant that describes the solubility of an electrolyte. And again, the pure
solid MX is not included in the expression since it has unit activity throughout the establishment of equilibrium.
Example 9.10.1:
What is the maximum solubility of CuS at 25 °C? (K sp = 1 × 10
−36
M
2
)
Solution
Yup – time for an ICE table.
2+ 2−
CuS Cu S
Initial 0 0
Change +x +x
Equilibrium x x
Solution
In this problem we need to consider the existence of S2-(aq) from the complete dissociation of the strong electrolyte NaS. An
ICE table will help, as usual.
2+ 2−
CuS Cu S
Initial 0 0.100 M
Change +x +x
Equilibrium x 0.100 M + x
Given the miniscule magnitude of the solubility product, x will be negligibly small compared to 0.100 MS the equilibrium
expression is
−36 2
1 × 10 M = x(0.100 M )
−35
1 × 10 M
9.10.1 https://chem.libretexts.org/@go/page/84558
The huge reduction in solubility is due to the common ion effect. The existence of sulfide in the solution due to sodium sulfide
greatly reduces the solutions capacity to support additional sulfide due to the dissociation of CuS.
This page titled 9.10: Solubility of Ionic Compounds is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
9.10.2 https://chem.libretexts.org/@go/page/84558
Welcome to the Chemistry Library. This Living Library is a principal hub of the LibreTexts project, which is a multi-institutional
collaborative venture to develop the next generation of open-access texts to improve postsecondary education at all levels of higher
learning. The LibreTexts approach is highly collaborative where an Open Access textbook environment is under constant revision
by students, faculty, and outside experts to supplant conventional paper-based books.
1 https://chem.libretexts.org/@go/page/84554
9.S: Chemical Equilibria (Summary)
Vocabulary and Concepts
common ion effect
conjugate base
degree of dissociation
dissociation of a weak acid
Dumas Bulb
Henderson-Hasselbalch equation
Le Chatlier’s principle
reaction quotient
solubility product
thermodynamic equilibrium constant
van’t Hoff equation
van’t Hoff plot
weak acid
References
1. Bandura, A. V., & Lvov, S. N. (2006). The Ionization Constant of Water over Wide Ranges of Temperature and Density. J.
Phys. Chem. Ref. Data, 35(1), 15-30.
2. Mack, E., & France, W. G. (1934). A Laboratory Manual of Elementary Physical Chemistry (2nd ed.). New York: D. Van
Nostrand Company, Inc.
3. Tro, N. J. (2014). Chemistry: a molecular approach (3rd ed.). Boston: Pearson.
This page titled 9.S: Chemical Equilibria (Summary) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
9.S.1 https://chem.libretexts.org/@go/page/84553
CHAPTER OVERVIEW
10: Electrochemistry
Selling an electric sports car creates an opportunity to fundamentally change the way
America drives.- Elon Musk
I've actually made a prediction that within 30 years a majority of new cars made in the
United States will be electric. And I don't mean hybrid, I mean fully electric.- Elon
Musk
Given the importance of energy production (and in particular, production from renewable sources) alluded to by Richard Smalley
in his address to the United States Congress (see Chapter 1), Elon Musk’s vision seems well-aligned with Smalley’s priority. The
generation and consumption of electrical energy and how it is harnessed to do work in the universe lends itself very nicely to
discussion within the framework of thermodynamics. In this chapter, we will use some of the tills we have developed to relate
electrochemical processes to thermodynamic variables, and to frame discussions of a few important topics.
10.1: Electricity
10.2: The connection to ΔG
10.3: Half Cells and Standard Reduction Potentials
10.4: Entropy of Electrochemical Cells
10.5: Concentration Cells
10.E: Electrochemistry (Exercises)
10.S: Electrochemistry (Summary)
This page titled 10: Electrochemistry is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick Fleming.
1
10.1: Electricity
Electricity has been known for some time. Ancient Egyptians, for example, referred to electric fish in the Nile River as early as
2750 BC (Moller & Kramer, 1991). In 1600, William Gilbert studied what would later be seen to be electrostatic attraction, by
creating static charges rubbing amber (Stewart, 2001). And Benjamin Franklin’s famous experiment (although it is actually
uncertain if he performed the experiment) of attaching a metal key to a kite string occurred in 1752, and showed that lightening is
an electrical phenomenon (Uman, 1987).
One of the biggest breakthroughs in the study of electricity as a chemical phenomenon was made by Alessandro Volta, who in 1799
showed that electricity could be generated by stacking copper and zinc disks submerged in sulfuric acid (Routledge, 1881). The
reactions that Volta produced in his voltaic pile included both oxidation and reduction processes that could be considered as half-
reactions. The half-reactions can be classified as oxidation (the loss of electrons) which happens at the anode and reduction (the
gain of electrons) which occurs at the cathode. Those half reactions were
2+ −
Zn → Z n + 2e
aanode
+ −
2H + 2e → H2
cathode
The propensity of zinc to oxidize coupled with that of hydrogen to reduce creates a potential energy difference between the
electrodes at which these processes occur. And like any potential energy difference, it can create a force which can be used to do
work. In this case, the work is that of pushing electrons through a circuit. The work of such a process can be calculated by
integrating
dwe − −E dQ
where E is the potential energy difference, and dQ is an infinitesimal amount of charge carried through the circuit. The
infinitesimal amount of charge carried through the circuit can be expressed as
dQ = e dN
we = −e E ∫ dN = −N e E
o
But since the number of electrons carried through a circuit is an enormous number, it would be far more convenient to express this
in terms of the number of moles of electrons carried through the circuit. Noting that the number of moles (n ) is given by
N
n =
NA
F = NA e = 96484 C
where F is Faraday's constant and has the magnitude of one Faraday (or the total charge carried by one mole of electrons.) The
Faraday is named after Michael Faraday (1791-1867) (Doc, 2014), a British physicist who is credited with inventing the electric
motor, among other accomplishments.
Putting the pieces together, the total electrical work accomplished by pushing n moles of electrons through a circuit with a potential
difference E , is
we = −nF E
This page titled 10.1: Electricity is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick Fleming.
10.1.1 https://chem.libretexts.org/@go/page/84351
10.2: The connection to ΔG
Recall that in addition to being used as a criterion for spontaneity, ΔG also indicated the maximum amount of non p-V work a
system could produce at constant temperature and pressure. And since we is non p-V work, it seems like a natural fit that
ΔG = −nF E
If all of the reactants and products in the electrochemical cell are in their standard states, it follows that
o o
ΔG = −nF E
where E is the standard cell potential. Noting that the molar Gibbs function change can be expressed in terms of the reaction
o
quotient Q by
o
ΔG = ΔG + RT ln Q
it follows that
o
−nF E = −nF E + RT ln Q
Dividing by – nF yields
o
RT
E =E − ln Q
nF
which is the Nernst equation. This relationship allows one to calculate the cell potential of a electrochemical cell as a function of
the specific activities of the reactants and products. In the Nernst equation, n is the number of electrons transferred per reaction
equivalent. For the specific reaction harnessed by Volta in his original battery, Eo = 0.763 V (at 25 oC) and n = 2 . So if the Zn2+
and H+ ions are at a concentration that gives them unit activity, and the H2 gas is at a partial pressure that gives it unit fugacity:
RT
E = 0.763 V − ln(1) = 0/763
nF
This page titled 10.2: The connection to ΔG is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
10.2.1 https://chem.libretexts.org/@go/page/84352
10.3: Half Cells and Standard Reduction Potentials
Much like G itself, E can only be measured as a difference, so a convention is used to set a zero to the scale. Toward this end,
convention sets the reduction potential of the standard hydrogen electrode (SHE) to 0.00 V.
2 + −
Zn → Zn +2 e
wtih Eo
ox = 0.763 V
+ −
2H +2 e → H
2
with Ered
o
= 0.000 V
+
Both H and H2 need to have unit activity (or fugacity), which if the solution and gas behave ideally means a concentration of 1 M
and a pressure of 1 bar.
Electrochemical Cells
Standard reduction potentials can be measured relative to the convention of setting the reduction potential of the Standard
Hydrogen Electrode (SHE) to zero. A number of values are shown in Table P1.
Fe and Cu
2 +
Solution
The species with the standard reduction potential (Table P1) will force the other to oxidize.
From the table,
2 + −
Cu +2 e → Cu
with 0.337 V
10.3.1 https://chem.libretexts.org/@go/page/84353
2 + −
Fe +2 e → Fe
with −0.440 V
So the iron half-reaction will flip (so that iron is oxidizing) and the spontaneous reaction under standard conditions will be
2+ 2+
Cu + F e → Cu + F e
with E o
= 0.777 V
Before calculating the cell potential, we should review a few definitions. The anode half reaction, which is defined by the half-
reaction in which oxidation °Ccurs, is
2 + −
Cu(s) → Cu (aq) + 2 e
And the cathode half-reaction, defined as the half-reaction in which reduction takes place, is
+ −
Ag (aq) + e → Ag(s)
Using standard cell notation, the conditions (such as the concentrations of the ions in solution) can be represented. In the standard
cell notation, the anode is on the left-hand side, and the cathode on the right. The two are typically separated by a salt bridge,
which is designated by a double vertical line. A single vertical line indicates a phase boundary. Hence for the reaction above, if the
silver ions are at a concentration of 0.500 M, and the copper (II) ions are at a concentration of 0.100 M, the standard cell notation
would be
Solution
In order to calculate the cell potential (E ), the standard cell potential must first be obtained. The standard cell potential at 25
°C is given by
o o
Ecell = E −E
cathode anode
= 0.799 V − 0.337 V
= 0.462 V
And for a cell at non-standard conditions, such as those indicated above, the Nernst equation can be used to calculate the cell
potential. At 25 °C, The cell potential is given by
2+
RT [C u ]
o
Ecell = E − ln( )
cell +
nF [Ag ]
E = 0.483 V
10.3.2 https://chem.libretexts.org/@go/page/84353
Example 10.3.3: Cell Potential under Non-Standard Conditions
Solution
We will use the Nernst equation. First, we need to determine E . Using Table P1, it is apparent that
o
2 + −
Cu 2e → Cu
o
E = 0.337 V
2 + −
Ni +2 e → Ni
with E o
= −0.250 V
So copper, having the larger reduction potential will be the cathode half-reaction while forcing nickel to oxidize, making it the
anode. So Eo for the cell will be given by
o o
Ecell = E −E
cathode anode
= 0.337 V − (−0.250 V )
= 0.587 V
= 0.566 V
10.3.3 https://chem.libretexts.org/@go/page/84353
This page titled 10.3: Half Cells and Standard Reduction Potentials is shared under a CC BY-NC-SA 4.0 license and was authored, remixed,
and/or curated by Patrick Fleming.
10.3.4 https://chem.libretexts.org/@go/page/84353
10.4: Entropy of Electrochemical Cells
The Gibbs function is related to entropy through its temperature dependence
∂ΔG
( ) = −ΔS
∂T
p
o
∂E
nF ( ) = ΔS (10.4.1)
∂T
p
Consider the following data for the Daniel cell (Buckbeei, Surdzial, & Metz, 1969) which is defined by the following reaction
2+ 2+
Zn(s) + C u (aq) ⇌ Z n (aq) + C u(s)
T (°C) 0 10 20 25 30 40
is easily established.
Figure 10.4.1 : Temperature dependence of the cell potential for a Daniel cell.
The quadratic fit to the data results in
o
∂E −6
V −4
V
( ) = 3.8576 × 10 (T ) − 6.3810 × 10
2
∂T °C °C
p
So, at 25 °C,
o
∂E
−4
( ) = −54166 × 10 V /K
∂T p
noting that K can be substituted for °C since in difference they have the same magnitude. So the entropy change is calculated
(Equation 10.4.1) is
o
∂E −4
ΔS = nF ( ) = (2 mol)(95484 C /mol)(−5.4166 × 10 V /K)
∂T
p
Because
1 C ×1 V = 1 J
10.4.1 https://chem.libretexts.org/@go/page/84354
ΔS = −104.5 J/(mol K).
It is the negative entropy change that leads to an increase in standard cell potential at lower temperatures. For a reaction such as
+ 2+
P b(s) + 2 H (aq) → P b (aq) + H2 (g)
which has a large increase in entropy (due to the production of a gas-phase product), the standard cell potential decreases with
decreasing temperature. As this is the reaction used in most car batteries, it explains why it can be difficult to start ones car on a
very cold winter morning. The topic of temperature dependence of several standard cell potentials is reported and discussed by
Bratsch (Bratsch, 1989).
This page titled 10.4: Entropy of Electrochemical Cells is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
10.4.2 https://chem.libretexts.org/@go/page/84354
10.5: Concentration Cells
The generation of an electrostatic potential difference is dependent on the creation of a difference in chemical potential between
two half-cells. One important manner in which this can be created is by creating a concentration difference. Using the Nernst
equation, the potential difference for a concentration cell (one in which both half-cells involve the same half-reaction) can be
expressed
RT [oxdizing]
E =− ln
nF [reducing]
Example 10.5.1
Calculate the cell potential (at 25 °C) for the concentration cell defined by
2+ 2+
C u(s)|C u (aq, 0.00100 M )||C u (aq, 0.100 M )|C u(s)
Solution
Since the oxidation and reduction half-reactions are the same,
o
E =0V
cell
= 0.059 V
This page titled 10.5: Concentration Cells is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
10.5.1 https://chem.libretexts.org/@go/page/84750
Welcome to the Chemistry Library. This Living Library is a principal hub of the LibreTexts project, which is a multi-institutional
collaborative venture to develop the next generation of open-access texts to improve postsecondary education at all levels of higher
learning. The LibreTexts approach is highly collaborative where an Open Access textbook environment is under constant revision
by students, faculty, and outside experts to supplant conventional paper-based books.
1 https://chem.libretexts.org/@go/page/84356
10.S: Electrochemistry (Summary)
References
1. Bratsch, S. G. (1989). Standard Electrode Potentialsand Temperature Coefficients in Water at 298.15 K. J. Phys. Chem. Ref.
Data, 18(1), 1-21.
2. Buckbeei, S. B., Surdzial, R. E., & Metz, C. R. (1969). Temperature Dependence of E° for the Daniel] Cell. Proceedings of the
Indiana Academy of Science, 79, pp. 123-128.
3. Doc, T. (2014, Novmber 24). Michael Faraday. Retrieved April 19, 2016, from famousscientists.org:
http://www.famousscientists.org/michael-faraday/
4. Moller, P., & Kramer, B. (1991). Review: Electric Fish. BioScience, 41(11), 794-796.
5. Routledge, R. (1881). A Popular History of Science (2nd ed.). G. Routledge and Sons.
6. Stewart, J. (2001). Intermediate Electromagnetic Theory. World Scientific.
7. Uman, M. (1987). All About Lightening. Dover.
Learning Objectives
This page titled 10.S: Electrochemistry (Summary) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
10.S.1 https://chem.libretexts.org/@go/page/84355
CHAPTER OVERVIEW
This page titled 11: Chemical Kinetics I is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
1
11.1: Reaction Rate
The rate of a chemical reaction (or the reaction rate) can be defined by the time needed for a change in concentration to occur. But
there is a problem in that this allows for the definition to be made based on concentration changes for either the reactants or the
products. Plus, due to stoichiometric concerns, the rates at which the concentrations are generally different! Toward this end, the
following convention is used.
For a general reaction
aA + bB → cC + dD
Example 11.1.1:
Under a certain set of conditions, the rate of the reaction
N2 + 3 H2 → 2N H3
Solution
Due to the stoichiometry of the reaction,
d[ N2 ] 1 d[ H2 ] 1 d[N H3 ]
rate = − =− =+
dt 3 dt 2 dt
so
d[ N2 ]
−4
= −6.0 × 10 M /s
dt
d[ H2 ]
−4
= −2.0 × 10 M /s
dt
d[N H3 ]
−4
= 3.0 × 10 M /s
dt
Note: The time derivatives for the reactants are negative because the reactant concentrations are decreasing, and those of
products are positive since the concentrations of products increase as the reaction progresses.
This page titled 11.1: Reaction Rate is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick Fleming.
11.1.1 https://chem.libretexts.org/@go/page/84358
11.2: Measuring Reaction Rates
There are several methods that can be used to measure chemical reactions rates. A common method is to use spectrophotometry to
monitor the concentration of a species that will absorb light. If it is possible, it is preferable to measure the appearance of a product
rather than the disappearance of a reactant, due to the low background interference of the measurement. However, high-quality
kinetic data can be obtained either way.
Some methods depend on measuring the initial rate of a reaction, which can be subject to a great deal of experimental uncertainty
due to fluctuations in instrumentation or conditions. Other methods require a broad range of time and concentration data. These
methods tend to produce more reliable results as they can make use of the broad range of data to smooth over random fluctuations
that may affect measurements. Both approaches (initial rates and full concentration profile data methods) will be discussed below.
This page titled 11.2: Measuring Reaction Rates is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
11.2.1 https://chem.libretexts.org/@go/page/84359
11.3: Rate Laws
A rate law is any mathematical relationship that relates the concentration of a reactant or product in a chemical reaction to time.
Rate laws can be expressed in either derivative (or ratio, for finite time intervals) or integrated form. One of the more common
general forms a rate law for the reaction
A + B → products
may take is
α β
rate = k[A] [B]
where k , α , and β are experimentally determined values. However, a rate law can take many different forms, some of which can be
quite intricate and complex. The powers α and β need not be integers. For example
α 1/2
rate = k[A] [B] (11.3.1)
is a rate law that is observed for some reactions. Sometimes, the concentrations of products must be included.
1/2
k[A] [B]
rate =
[P ]
In some cases, the concentration for a catalyst or enzyme is important. For example, many enzyme mitigated reactions in biological
systems follow the Michaelis-Menten rate law, which is of the form
Vmax [S]
rate =
Km + [S]
where V max and K M are factors that are determined experimentally, and [S] is the concentration of the substrate in the reaction.
Order
For those cases where the rate law can be expressed in the form
α β γ
rate = k[A] [B] [C ]
where A , B , and C are reactants (or products or catalysts, etc.) involved in the reaction, the reaction is said to be of α order in A ,
β order in B , and γ order in C . The reaction is said to be α + β + γ order overall. Some examples are shown in the following
table:
Table 11.3.1: Example Rate Laws
Rate law Order with respect to A Order with respect to B Order with respect to C Overall order
rate = k 0 0 0 0
rate = k[A] 0 0 0 1
rate = k[A]
2
0 0 0 2
rate = k[A][B] 0 1 0 2
2
rate = k[A ] [B] 0 1 0 3
rate = k[A][B][C] 0 1 1 3
Reaction orders can also be fractional such as for Equation 11.3.1 which is 1st order in A , and half order in B . The order can also
be negative such as
[A]
rate = k
[B]
which is 1st order in A, and -1 order in B. In this case, an build-up of the concentration of B will retard (slow) the reaction.
In all cases, the order of the reaction with respect to a specific reactant or product (or catalyst, or whatever) must be determined
experimentally. As a general rule, the stoichiometry cannot be used to predict the form of the rate law. However, the rate law can be
11.3.1 https://chem.libretexts.org/@go/page/84360
used to gain some insight into the possible pathways by which the reaction can proceed. That is the topic of Chapter 12. For now
we will focus on three useful methods that are commonly used in chemistry to determine the rate law for a reaction from
experimental data.
Empirical Methods
Perhaps the simplest of the methods to be used are the empirical methods, which rely on the qualitative interpretation of a graphical
representation of the concentration vs time profile. In these methods, some function of concentration is plotted as a function of
time, and the result is examined for a linear relationship. For the following examples, consider a reaction of the form
A + B → products
in which A is one of the reactants. In order to employ these empirical methods, one must generate the forms of the integrated rate
laws.
This page titled 11.3: Rate Laws is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick Fleming.
11.3.2 https://chem.libretexts.org/@go/page/84360
11.4: 0th order Rate Law
If the reaction follows a zeroth order rate law, it can be expressed in terms of the time-rate of change of [A] (which will be negative
since A is a reactant):
d[A]
− =k
dt
In this case, it is straightforward to separate the variables. Placing time variables on the right and [A] on the left
d[A] = −k dt
In this form, it is easy to integrate. If the concentration of A is [A]0 at time t = 0, and the concentration of A is [A] at some arbitrary
time later, the form of the integral is
[A] t
∫ d[A] = −k ∫ dt
[A] to
o
which yields
or
[A] = [A]o − kt
This suggests that a plot of concentration as a function of time will produce a straight line, the slope of which is –k, and the
intercept of which is [A]0. If such a plot is linear, then the data are consistent with 0th order kinetics. If they are not, other
possibilities must be considered.
This page titled 11.4: 0th order Rate Law is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
11.4.1 https://chem.libretexts.org/@go/page/84361
11.5: 1st order rate law
A first order rate law would take the form
d[A]
= k[A]
dt
Again, separating the variables by placing all of the concentration terms on the left and all of the time terms on the right yields
d[A]
= −k dt
[A]
Noting that
dx
= d(ln x)
x
ln[A] − ln[A]o = kt
or
ln[A] = ln[A]o − kt (11.5.1)
This form implies that a plot of the natural logarithm of the concentration is a linear function of the time. And so a plot of ln[A] as
a function of time should produce a linear plot, the slope of which is -k, and the intercept of which is ln[A]0.
Example 11.5.1:
Consider the following kinetic data. Use a graph to demonstrate that the data are consistent with first order kinetics. Also, if the
data are first order, determine the value of the rate constant for the reaction.
[A] (M) 0.873 0.752 0.648 0.414 0.196 0.093 0.044 0.021 0.010
Solution
The plot looks as follows:
11.5.1 https://chem.libretexts.org/@go/page/84362
From this plot, it can be seen that the rate constant is 0.0149 s-1. The concentration at time t =0 can also be inferred from the
intercept.
It should also be noted that the integrated rate law (Equation 11.5.1) can be expressed in exponential form:
−kt
[A] = [A]o e
Because of this functional form, 1st order kinetics are sometimes referred to as exponential decay kinetics. Many processes,
including radioactive decay of nuclides follow this type of rate law.
This page titled 11.5: 1st order rate law is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
11.5.2 https://chem.libretexts.org/@go/page/84362
11.6: 2nd order Rate Laws
If the reaction follows a second order rate law, the some methodology can be employed. The rate can be written as
d[A]
2
− = k[A] (11.6.1)
dt
The separation of concentration and time terms (this time keeping the negative sign on the left for convenience) yields
d[A]
− = k dt
2
[A]
or
1 1
= + kt
[A] [A]o
And so a plot of 1/[A] as a function of time should produce a linear plot, the slope of which is k , and the intercept of which is
1/[A] .
0
Other 2nd order rate laws are a little bit trickier to integrate, as the integration depends on the actual stoichiometry of the reaction
being investigated. For example, for a reaction of the type
A+B → P
and
d[B]
− = k[A][B]
dt
the integration will depend on the decrease of [A] and [B] (which will be related by the stoichiometry) which can be expressed in
terms the concentration of the product [P].
[A] = [A]o – [P ] (11.6.3)
and
[B] = [B]o – [P ] (11.6.4)
The concentration dependence on A and B can then be eliminated if the rate law is expressed in terms of the production of the
product.
d[P ]
= k[A][B] (11.6.5)
dt
11.6.1 https://chem.libretexts.org/@go/page/84363
Substituting the relationships for [A] and [B] (Equations 11.6.3 and 11.6.4) into the rate law expression (Equation 11.6.5) yields
d[P ]
= k([A]o – [P ])([B] = [B]o – [P ]) (11.6.6)
dt
Noting that at time t = 0 , [P ] = 0 , the integrated form of the rate law can be generated by solving the integral
[A] t
d[P ]
∫ =∫ k dt
[A] ([A]o – [P ])([B]o – [P ]) t=0
o
1 [B]0 − [P ] 1 [B]0
ln( )− ln( ) =kt (11.6.7)
[B]0 − [A]0 [A]0 − [P ] [B]0 − [A]0 [A]0
Substituting Equations 11.6.3 and 11.6.4 into Equation 11.6.7 and simplifying (combining the natural logarithm terms) yields
1 [B][A]o
ln( ) = kt
[B]0 − [A]0 [A][B]o
For this rate law, a plot of ln([B]/[A]) as a function of time will produce a straight line, the slope of which is
In the limit at [A] 0 = [B]0 , then [A] = [B] at all times, due to the stoichiometry of the reaction. As such, the rate law becomes
2
rate = k[A]
and integrate direct like in Equation 11.6.1 and the integrated rate law is (as before)
1 1
= + kt
[A] [A]o
Consider the following kinetic data. Use a graph to demonstrate that the data are consistent with second order kinetics. Also, if
the data are second order, determine the value of the rate constant for the reaction.
Solution
The plot looks as follows:
11.6.2 https://chem.libretexts.org/@go/page/84363
From this plot, it can be seen that the rate constant is 0.2658 M-1 s-1. The concentration at time t = 0 can also be inferred from
the intercept.
[1] This integral form can be generated by using the method of partial fractions. See (House, 2007) for a full derivation.
This page titled 11.6: 2nd order Rate Laws is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
11.6.3 https://chem.libretexts.org/@go/page/84363
11.7: The Method of Initial Rates
The method of initial rates is a commonly used technique for deriving rate laws. As the name implies, the method involves
measuring the initial rate of a reaction. The measurement is repeated for several sets of initial concentration conditions to see how
the reaction rate varies. This might be accomplished by determining the time needed to exhaust a particular amount of a reactant
(preferably one on which the reaction rate does not depend!) A typical set of data for a reaction
A + B → products
The analysis of this data involves taking the ratios of rates measured where one of the concentrations does not change. For
example, assuming a rate law of the form
α β
rate = k[A] [B] (11.7.1)
this simplifies to
α
1 1
=( )
2 2
So clearly, \(\alpha = 1\) and the reaction is 1st order in A . Taking the ratio using runs 2 and 3 yields
α β
k (0.02 M ) (0.01 M )
0.0694 M /s
=
0.2776 M /s α β
k (0.02 M ) (0.02 M )
This simplifies to
β
1 1
=( ) (11.7.2)
4 2
By inspection, one can conclude that \(\beta = 2\), and that the reaction is second order in B. But if it is not so clear (as it might not
be if the concentration is not incremented by a factor of 2), the value of \(\beta\) can be determined by taking the natural logarithm
of both sides of the Equation 11.7.2.
β
1 1
ln = ln ( )
4 2
1
= β ln( )
2
11.7.1 https://chem.libretexts.org/@go/page/84581
1 1
ln( ) ln( )
4 2
=β
1 1
ln( ) ln( )
2 2
or
−1.3863
β = =2
−0.69315
And is 1st order in A, 2nd order in B, and 3rd order overall. The rate constant can then be evaluated by substituting one of the runs
into the rate law (or using all of the data and taking an average). Arbitrarily selecting the first run for this,
2
0.0347 M /s = k(0.01 M /s)(0.01 M /s)
It is useful to note that the units on k are consistent with a 3rd order rate law.
This page titled 11.7: The Method of Initial Rates is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
11.7.2 https://chem.libretexts.org/@go/page/84581
11.8: The Method of Half-Lives
Another method for determining the order of a reaction is to examine the behavior of the half-life as the reaction progresses. The
half-life can be defined as the time it takes for the concentration of a reactant to fall to half of its original value. The method of
half-lives involved measuring the half-life’s dependence on concentration. The expected behavior can be predicted using the
integrated rate laws we derived earlier.
Using the definition of the half-life, at time t 1/2 the concentration [A] drops to half of its original value, [A] . 0
1
[A] = [A]o
2
at t = t1/2 .
So if the reaction is 0th order in A , after one half-life
1
[A]o = [A]o − kt1/2
2
So as the original concentration is decreased, the half-life of a 0th order reaction will also decrease.
Similarly, for a first order reaction,
1 −kt1/2
[A]o = [A]o e
2
It is because the half-life of a 1st order reaction is independent of concentration that it is oftentimes used to describe the rate of first
order processes (such as radioactive decay.)
For a 2nd order reaction, the half-life can be expressed based on the integrated rate law.
1 1
= + kt1/2
1 [A]o
[A]o
2
In the case of a second order reaction, the half-life increases with decreasing initial concentration.
Table 11.8.1 : Calculated half lives for Reactions following simple Rate Laws
Order Half-life Behavior
For reactions in which the rate law depends on the concentration of more than one species, the half-life can take a much more
complex form that may depend on the initial concentrations of multiple reactants, or for that matter, products!
11.8.1 https://chem.libretexts.org/@go/page/84582
Example 11.8.1: Radiocarbon Dating
Carbon-14 decays into nitrogen-14 with first order kinetics and with a half-life of 5730 years.
14 14
C → N
What is the rate constant for the decay process? What percentage of carbon-14 will remain after a biological sample has
stopped ingesting carbon-14 for 1482 years?
Solution
The rate constant is fairly easy to calculate:
ln 2 ln 2
−4 −1
t1/2 = = = 1.21 × 10 yr
k 5730 yr
Now the integrated rate law can be used to solve the second part of the problem.
14 14 −kt
[ C] = [ C]o e
so
14
[ C] −4 −1
−(1.21×10 yr )(1482 ys)
=e = 0.836
14
[ C]o
Example 11.8.2:
Based on the following concentration data as a function of time, determine the behavior of the half-life as the reaction
progresses. Use this information to determine if the following reaction is 0th order, 1st order, or 2nd order in A. Also, use the
data to estimate the rate constant for the reaction.
0 1.200
10 0.800
20 0.600
30 0.480
40 0.400
50 0.343
60 0.300
70 0.267
80 0.240
90 0.218
100 0.200
Solution
11.8.2 https://chem.libretexts.org/@go/page/84582
If the original concentration is taken as 1.200 M, half of the original concentration is 0.600 M. The reaction takes 20 seconds to
reduce the concentration to half of its original value. If the original concentration is taken as 0.800 M, it clearly takes 30
seconds for the concentration to reach half of that value. Based on this methodology, the following table is easy to generate:
t
1/2
(s) 20 30 40 60
1
= (11.8.2)
(0.8 M )(30 s)
−1 −1
= 0.0417 M s (11.8.3)
This page titled 11.8: The Method of Half-Lives is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
11.8.3 https://chem.libretexts.org/@go/page/84582
11.9: Temperature Dependence
In general, increases in temperature increase the rates of chemical reactions. It is easy to see why, since most chemical reactions
depend on molecular collisions. And as we discussed in Chapter 2, the frequency with which molecules collide increases with
increased temperature. But also, the kinetic energy of the molecules increases, which should increase the probability that a collision
event will lead to a reaction. An empirical model was proposed by Arrhenius to account for this phenomenon. The Arrhenius
model (Arrhenius, 1889) can be expressed as
−Ea /RT
k = Ae
Although the model is empirical, some of the parameters can be interpreted in terms of the energy profile of the reaction. E , for a
example, is the activation energy, which represents the energy barrier that must be overcome in a collision to lead to a reaction.
Figure 11.9.1: Reaction profile for an elementary sttep over an activated barrier of height E .
a
If the rate constant for a reaction is measure at two temperatures, the activation energy can be determined by taking the ratio. This
leads to the following expression for the Arrhenius model:
k1 Ea 1 1
ln( ) =− ( − ) (11.9.1)
k2 R T2 T1
Example 11.9.1:
For a given reaction, the rate constant doubles when the temperature is increased form 25 °C to 35 °C. What is the Arrhenius
activation energy for this reaction?
Solution
The energy of activation can be calculated from the Arrhenius Equation (Equation 11.9.1).
2k1 Ea 1 1
ln( ) =− ( − )
k1 J 308 K 298 K
8.314
mol K
Ea = 52.9 kJ/mol
Preferably, however, the rate constant is measured at several temperatures, and then the activation energy can be determined using
all of the measurements, by fitting them to the expression
Ea
ln(k) = − + ln(A)
RT
11.9.1 https://chem.libretexts.org/@go/page/84583
This can be done graphically by plotting the natural logarithm of the rate constant as a function of 1/T (with the temperature
measured in K). The result should be a straight line (for a well-behaved reaction!) with a slope of – E /R.a
There are some theoretical models (such as collision theory and transition state theory) which suggest the form of the Arrhenius
model, but the model itself is purely empirical. A general feature, however, of the theoretical approaches is to interpret the
activation energy as an energy barrier which a reaction must overcome in order to lead to a chemical reaction.
This page titled 11.9: Temperature Dependence is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
11.9.2 https://chem.libretexts.org/@go/page/84583
11.10: Collision Theory
Collision Theory was first introduced in the 1910s by Max Trautz (Trautz, 1916) and William Lewis (Lewis, 1918) to try to
account for the magnitudes of rate constants in terms of the frequency of molecular collisions, the collisional energy, and the
relative orientations of the molecules involved in the collision.
The rate of a reaction, according to collision theory, can be expressed as
rate = Zab F (11.10.1)
where Z AB is the frequency of collisions between the molecules A and B involved in the reaction, and F is the fraction of those
collisions that will lead to a reaction. The factor F has two important contributors, the energy of the collision and the orientation of
the molecules when they collide. The first term, Z , can be taken from the kinetic molecular theory discussed in Chapter 2.
AB
1/2
8 kB T
ZAB = ( ) σAB [A][B] (11.10.2)
πμ
Where the first term is the average relative velocity in which μ is the reduced mass of the A-B collisional system, σAB is the
collisional cross section, and [A] and [B] are the concentrations of A and B .
The factor F depends on the activation energy. Assuming a Boltzmann (or Boltzmann-like) distribution of energies, the fraction of
molecular collisions that will have enough energy to overcome the activation barrier is given by
−Ea /RT
F =e (11.10.3)
Combining Equations 11.10.2 and 11.10.3, the rate of the reaction (Equation 11.10.1) is predicted by
1/2
8 kB T
−Ea /RT
rate = ( ) σAB e [A][B]
πμ
rate = k[A][B]
It should be noted that collision theory appears to apply only to bimolecular reactions, since it takes two molecules to collide. But
there are many reactions that have first order rate laws, but are initiated by bimolecular steps in the mechanisms. (Reaction
mechanisms will form a large part of the discussion in Chapter 12.) Consider as an example, the decomposition of N O , which 2 5
2 N2 O5 → 4N O2 + O2
Under a certain set of conditions, the following concentrations are observed as a function of time.
11.10.1 https://chem.libretexts.org/@go/page/84593
time (s) [ N2 O5 ] (M) [N O2 ] (M) [ O2 ] (M)
So how can collision theory be used to understand the rate constant? As it turns out, the mechanism for the reaction involves a
bimolecular initiation step.
where N O is an energetically activated form of N O which can either relax to reform N O or decompose to form the products
2
∗
5 2 5 2 5
of the reaction. Because the initiation step is bimolecular, collision theory can be used to understand the rate law, but because the
product of the unimolecular step undergoes slow conversion to products unimolecularly, the overall rate is observed to be first
order in N O . The analysis of reaction mechanisms, and reconciliation with observed rate laws, form the subjects of Chapter 12.
2 5
This page titled 11.10: Collision Theory is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
11.10.2 https://chem.libretexts.org/@go/page/84593
11.11: Transition State Theory
Transition state theory was proposed in 1935 by Henry Erying, and further developed by Merrideth G. Evans and Michael
Polanyi (Laidler & King, 1983), as another means of accounting for chemical reaction rates. It is based on the idea that a molecular
collision that leads to reaction must pass through an intermediate state known as the transition state. For example, the reaction
A + BC → AB + C
would have an intermediate (ABC ) where the B − C bond is partially broken, and the A − B bond is partially formed.
‡
A + B − C → (A − B − C ) → A−B+C
So the reaction is mediated by the formation of an activated complex (denoted with the double-dagger symbol ‡ ) and the
decomposition of that complex into the reaction products. Using this theory, the rate of reaction can be expressed as the product of
two factors
Using the relationship from Chapter 9 for the equilibrium constant, K can be expressed in terms of the free energy of formation of
‡
and the remaining task is to derive an expression for the frequency factor. If the frequency is taken to be equal to the vibrational
frequency for the vibration of the bond being broken in the activated complex in order to form the reaction products, it can be
expressed in terms of the energy of the oscillation of the bond as the complex vibrates.
E = hν = kB T
or
kB T
ν =
h
k = e
h
An alternative description gives the transition state formation equilibrium constant in terms of the partition functions describing the
reactants and the transition state:
‡
Q ‡
‡ −Δ G /RT
K = e
‡ ‡
Q Q
A BC
11.11.1 https://chem.libretexts.org/@go/page/84594
where Qi is the partition function describing the ith species. If the partition function of the transition state is expressed as a product
of the partition function excluding and contribution from the vibration leading to the bond cleavage that forms the products and the
partition function of that specific vibrational mode
′
‡ ‡ ‡
Q =Q qv
‡
In this case, q can be expressed by
v
‡ 1 kB T
qv = ≈
‡
−hν /RT ‡
1 −e hν
‡
kB T Q ‡
−Δ G /RT
k = e
‡ ‡
h
Q Q
A BC
which looks very much like the Arrhenius equation proposed quite a few years earlier! Thus, if one understands the vibrational
dynamics of the activated complex, and can calculate the partition functions describing the reactants and the transition state, one
can, at least in theory, predict the rate constant for the reaction. In the next chapter, we will take a look at how kinetics studies can
shed some light on chemical reaction mechanisms.
This page titled 11.11: Transition State Theory is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
11.11.2 https://chem.libretexts.org/@go/page/84594
Welcome to the Chemistry Library. This Living Library is a principal hub of the LibreTexts project, which is a multi-institutional
collaborative venture to develop the next generation of open-access texts to improve postsecondary education at all levels of higher
learning. The LibreTexts approach is highly collaborative where an Open Access textbook environment is under constant revision
by students, faculty, and outside experts to supplant conventional paper-based books.
1 https://chem.libretexts.org/@go/page/84580
11.S: Chemical Kinetics I (Summary)
The results of the integration of these simple rate laws can be summarized in the following table.
ln[A] = ln[A ]o − kt
1 A → P
−kt
ln[A] vs. t
[[A] = [A]o e
1 1 1
A +A → P = + kt vs. t
[A] [A]o [A]
2
1 [B][A]o [B]
A +B → P ln( ) = kt ln( ) vs. t
[B ]0 − [A ]0 [A][B]o [A]
This page titled 11.S: Chemical Kinetics I (Summary) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
11.S.1 https://chem.libretexts.org/@go/page/84579
CHAPTER OVERVIEW
This page titled 12: Chemical Kinetics II is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
1
12.1: Reaction Mechanisms
A reaction mechanism is a set of elementary reactions steps, that when taken in aggregate define a chemical pathway that
connects reactants to products. An elementary reaction is one that proceeds by a single process, such a molecular (or atomic)
decomposition or a molecular collision. Typically, elementary reactions only come in unimolecular
A → products
and bimolecular
A + B → products
A + B + C → products
(involved the simultaneous collision of three atoms or molecules) but it is generally a pair of bimolecular steps acting in rapid
succession, the first forming an activated complex, and the second stabilizing that complex chemically or physically.
∗
A + B → AB
∗ ∗
AB + C → AB + C
The wonderful property of elementary reactions is that the molecularity defines the order of the rate law for the reaction step.
Example 12.1.1:
For the reaction
A+B → C
A+A−
→ A2
k1
A2 + B −
→ C +A
Solution
Adding both proposed reactions gives
2 A + A2 + B → A2 +C + A
Canceling those species that appear on both sides of the arrow leaves
A+B → C
which is the reaction, so the mechanism is at least stoichiometrically valid. However, it would still have to be consistent with
the observed kinetics for the reaction and account for any side-products that are observed.
This page titled 12.1: Reaction Mechanisms is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
12.1.1 https://chem.libretexts.org/@go/page/84365
12.2: Concentration Profiles for Some Simple Mechanisms
To illustrate how mechanisms may affect the concentration profile for a reaction, we can examine some simple mechanisms
A → B
In this type of reaction, one substance is simply converting into another. An example of this type of reaction might be the
isomerization of methylisocyanide to form acetonitrile (methylcyanide) (Redmon, Purvis, & Bartlett, 1978). If the reaction
mechanism consists of a single unimolecular step, which is characterized by the rate constant k :
1
k1
A−
→B
and
d[B]
= +k1 [A]
dt
It can be easily seen that the concentration of the reactant (A) decreases as time moves forward, and that of the product (B)
increases. This will continue until reactant A is depleted.
A ⇌ B
When the system can establish equilibrium, the rate of change of the concentration of A and B will depend on both the forward and
reverse reactions. If k1 is the rate constant that characterizes the forward reaction
k1
A−
→B
B−
−→A
and
d[B]
= +k1 [A] − k−1 [B]
dt
12.2.1 https://chem.libretexts.org/@go/page/84366
This profile is characterized by the fact that after a certain amount of time, the system achieves equilibrium and the concentrations
stop changing (even though the forward and reverse reactions are still taking place. This is the nature of a dynamic equilibrium
about which we speak off of the time in chemistry. The final concentrations of [A] and [B] once equilibrium is established will
depend on the ratio of k1 and k-
Since the rate of formation of A (from the reverse step) is equal to the rate of consumption of A (from the forward step, the overall
rate of change of the concentration of A is zero once equilibrium has been established. So it should be clear that
or
k1 [B]
=
k−1 [A]
A +C → B +C
Some reactions require a catalyst to mediate the conversion of reactants in to products. The definition of a catalyst is a species that
must be added (it is not formed as an intermediate) shows up in the mechanism (usually in a very early step) and this ends up as
part of the rate law, but is reformed later on so that it does not appear in the overall stoichiometry. If the reaction
A → B
A+C → B+C
In this case, C is acting as a catalyst to the reaction. The rate of change of the concentrations can be found by
[A]
= −k[A][C ]
dt
[B]
= k[A][C ]
dt
[C ]
= −k[A][C ] + k[A][C ] = 0
dt
12.2.2 https://chem.libretexts.org/@go/page/84366
This is a very simplified picture of a catalyzed reaction. Generally a catalyzed reaction will require at least two steps:
A + C → AC
AC → B + C
Later, we will see how the steady-state approximation actually predicts the above depicted concentration profile for the two-step
mechanism when AC is a short-lived species that can be treated as having a constant and small concentration.
A → B → C
Another important (and very common) mechanistic feature is the formation of an intermediate. This is a species that is formed in at
least one of the mechanism step, but does not appear in the overall stoichiometry for the reaction. This is different from a catalyst
which must be added to speed the reaction. A simple example of a reaction mechanism involving the formation of a catalyst is
k1
A−
→B
k2
B−
→C
In this case, C cannot form until an appreciable concentration of the intermediate B has been created by the first step of the
mechanism.
The rate of change of the concentrations of A , B , and C can be expressed
[A]
= −k1 [A]
dt
[B]
= k1 [A] − k2 [B]
dt
[C ]
= k2 [B]
dt
The concentration profile is then shown below. Notice the delay in the formation of C .
12.2.3 https://chem.libretexts.org/@go/page/84366
A ⇌ B → C
In many cases, the formation of an intermediate involves a reversible step. This step is sometimes referred to as a pre-equilibrium
step since it oftentimes will establish a near equilibrium while the reaction progresses. The result of combining a pre-equilibrium
with an intermediate produces a profile that shows features of both of the simpler mechanisms. An example of such a mechanism is
k−1
A ⇌ B
k1
k2
B−
→C
In this case, the rate of change for the concentrations of A , B , and C can be expressed by
[A]
= −k1 [A] + −k−1 [B]
dt
[B]
= k1 [A] − k−1 [B] − k2 [B]
dt
[C ]
= k2 [B]
dt
The concentration profile for this mechanism is shown below. Again, notice the delay in the production of the product C , due to the
requirement that the concentration of B be sufficiently high to allow the second step to occur with an appreciable rate.
A → B and A → C
There are many cases where a reactant can follow pathways to different products (or sometimes even the same products!), and
those pathways compete with one another. An example is the following simple mechanism:
k−1
A−
−→B
k2
A−
→C
12.2.4 https://chem.libretexts.org/@go/page/84366
In this case, the rate of change on concentrations can be expressed as
[A]
= −k1 [A] + −k2 [A]
dt
[B]
= +k1 [A]
dt
[C ]
= +k2 [A]
dt
Overall, the profile looks like two first order decompositions occurring at the same time, with the final concentration of the product
formed with the larger rate constant being favored.
One of the goals of studying chemical kinetics is to understand how to alter reaction condition to favor the production of desirable
reaction products. This can be accomplished by a number of means, such as alteration of concentrations, temperature, addition of
catalysts, etc. Understanding the basics will (hopefully) lead to a better understanding of how concentration profiles can be altered
by changing conditions.
This page titled 12.2: Concentration Profiles for Some Simple Mechanisms is shared under a CC BY-NC-SA 4.0 license and was authored,
remixed, and/or curated by Patrick Fleming.
12.2.5 https://chem.libretexts.org/@go/page/84366
12.3: The Connection between Reaction Mechanisms and Reaction Rate Laws
The great value of chemical kinetics is that it can give us insights into the actual reaction pathways (mechanisms) that reactants
take to form the products of reactions. Analyzing a reaction mechanism to determine the type of rate law that is consistent (or not
consistent) with the specific mechanism can give us significant insight. For example, the reaction
A+B → C
A+A− → A2
step 1
k2
A2 + B − →C
step 2
or
k1
∗
A− →A
step 1
k2
∗
A +B− →C
step 2
The first rate law will predict that the reaction should be second order in A , whereas the second mechanism predicts that it should
be first order in A (in the limit that the steady state approximation, discussed in the following sections, can be applied to A and
2
A ). Based on the observed rate law being first or second order in A, one can rule out one of the rate laws. Unfortunately, this kind
∗
of analysis cannot confirm a specific mechanism. Other evidence is needed to draw such conclusions, such as the spectroscopic
observation of a particular reaction intermediate that can only be formed by a specific mechanism.
In order analyze mechanisms and predict rate laws, we need to build a toolbox of methods and techniques that are useful in certain
limits. The next few sections will discuss this kind of analysis, specifically focusing on
the Rate Determining Step approximation,
the Steady State approximation, and
the Equilibrium approximation.
Each type of approximation is important in certain limits, and they are oftentimes used in conjunction with one another to predict
the final forms of rate laws.
This page titled 12.3: The Connection between Reaction Mechanisms and Reaction Rate Laws is shared under a CC BY-NC-SA 4.0 license and
was authored, remixed, and/or curated by Patrick Fleming.
12.3.1 https://chem.libretexts.org/@go/page/84367
12.4: The Rate Determining Step Approximation
The rate determining step approximation is one of the simplest approximations one can make to analyze a proposed mechanism
to deduce the rate law it predicts. Simply stated, the rate determining step approximation says that a mechanism can proceed no
faster than its slowest step. So, for example, if the reaction
A+B → C
A+A− → A2
slow
k2
A2 − → C +A
fast
the rate determining step approximation suggests that the rate (expressed in terms of the appearance of product C ) should be
determined by the slow initial step, and so the rate law will be
[C ]
2
= k1 [A]
dt
matching the order of the rate law to the molecularity of the slow step. Conversely, if the reaction mechanism is proposed as
k1
∗
A− →A
slow
k2
∗
A +B− →C
fast
the rate determining step approximation suggests that the rate of the reaction should be
[C ]
= k1 [A]
dt
again, with the order of the rate law matching the molecularity of the rate determining step.
This page titled 12.4: The Rate Determining Step Approximation is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by Patrick Fleming.
12.4.1 https://chem.libretexts.org/@go/page/84368
12.5: The Steady-State Approximation
One of the most commonly used and most attractive approximations is the steady state approximation. This approximation can be
applied to the rate of change of concentration of a highly reactive (short lived) intermediate that holds a constant value over a long
period of time. The advantage here is that for such an intermediate (I ),
d[I ]
=0
dt
So long as one can write an expression for the rate of change of the concentration of the intermediate I , the steady state
approximation allows one to solve for its constant concentration. For example, if the reaction
A+B → C (12.5.1)
A+A −
→ A2 (12.5.2)
k2
A2 + B −
→ C +A (12.5.3)
d[ A2 ]
2
= k1 [A] − k2 [ A2 ][B]
dt
d[ A2 ]
2
= k1 [A] − k2 [ A2 ][B] ≈ 0
dt
or
2
k1 [A]
[ A2 ] ≈
k2 [B]
So if the rate of the overall reaction is expressed as the rate of formation of the product C ,
d[C ]
= k2 [ A2 ][B]
dt
of
d[C ]
2
= k1 [A]
dt
k2
∗
A +B −
→C (12.5.5)
∗
[A ]
∗
= k1 [A] − k2 [ A ][B]
dt
12.5.1 https://chem.libretexts.org/@go/page/84369
k1 [A]
∗
[A ] ≈
k2 [B]
⎛ k1 [A] ⎞
= k2 [B] (12.5.7)
⎝ k2 [B] ⎠
or
d[C ]
= k1 [A]
dt
and the rate law is predicted to be first order in A . In this manner, the plausibility of either of the two reaction mechanisms is easily
deduced by comparing the predicted rate law to that which is observed. If the prediction cannot be reconciled with observation,
then the scientific method eliminates that mechanism from consideration.
This page titled 12.5: The Steady-State Approximation is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
12.5.2 https://chem.libretexts.org/@go/page/84369
12.6: The Equilibrium Approximation
In many cases, the formation of a reactive intermediate (or even a longer lived intermediate) involves a reversible step. This is the
case if the intermediate can decompose to reform reactants with a significant probability as well as moving on to form products. In
many cases, this will lead to a pre-equilibrium condition in which the equilibrium approximation can be applied. An example of a
reaction mechanism of this sort is
k−1
A+B ⇌ AB
k1
k2
AB −
→C
Given this mechanism, the application of the steady state approximation is cumbersome. However, if the initial step is assumed to
achieve equilibrium, an expression can be found for [AB]. In order to derive this expression, one assumes that the rate of the
forward reaction is equal to the rate of the reverse reaction for the initial step in the mechanism.
or
k1 [A][B]
= [AB]
k−1
This expression can be substituted into an expression for the rate of formation of the product C :
d[C ]
= k2 [AB]
dt
or
d[C ] k2 k1
= [A][B]
dt k−1
Which predicts a reaction rate law that is first order in A , first order in B , and second order overall.
Example 12.6.1:
Given the following mechanism, apply the equilibrium approximation to the first step to predict the rate law suggested by the
mechanism.
k−1
A+A ⇌ A2
k1
k2
A2 + B −
→ C +A
Solution
If the equilibrium approximation is valid for the first step,
2
k1 [A] = k−1 [ A2 ]
or
2
k1 [A]
≈ [ A2 ]
k−1
Plugging this into the rate equation for the second step
d[C ]
= k2 [ A2 ][B]
dt
12.6.1 https://chem.libretexts.org/@go/page/84370
yields
d[C ] k2 k1
2
= [A] [B]
dt k−1
which is second order in A , first order in B and third order over all, and in which the effective rate constant (k is ′
k2 k1
′
k = .
k−1
Sometimes, the equilibrium approximation can suggest rate laws that have negative orders with respect to certain species. For
example, consider the following reaction
A + 2B → 2C
A+B ⇌ I +C
k1
k2
I +B−
→C
in which I is an intermediate. Applying the equilibrium approximation to the first step yields
or
k1 [A][B]
≈ [I ]
k−1 [C ]
Substituting this into an expression for the rate of formation of C , one sees
d[C ]
= k2 [I ][B]
dt
or
d[C ] k1 [A][B] k2 k1 [A][B]
= [B] =
dt k−1 [C ] k−1 [C ]
which is first order in A , second order in B , negative one order in C , and second order overall. Also,
′
k2 k1
k = .
k−1
In this case, the negative order in C means that a buildup of compound C will cause the reaction to slow. These sort of rate laws
are not uncommon for reactions with a reversible initial step that forms some of the eventual reaction product.
This page titled 12.6: The Equilibrium Approximation is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
12.6.2 https://chem.libretexts.org/@go/page/84370
12.7: The Lindemann Mechanism
The Lindemann mechanism (Lindemann, Arrhenius, Langmuir, Dhar, Perrin, & Lewis, 1922) is a useful one to demonstrate some
of the techniques we use for relating chemical mechanisms to rate laws. In this mechanism, a reactant is collisionally activated to a
highly energetic form that can then go on to react to form products.
k−1
∗
A+A ⇌ A
k1
k2
∗
A −
→P
∗
d[ A ]
2 ∗ ∗
= k1 [A] − k−1 [ A ][A] − k2 [ A ] ≈ 0
dt
2
k1 [A]
∗
A ] =
k−1 [A] + k2
Substituting this into an expression for the rate of the production of the product P
d[P ]
∗
= k2 [ A ]
dt
yields
2
d[P ] k2 k1 [A]
=
dt k−1 [A] + k2
This will happen if the second step is very slow (and is the rate determining step), such that the reverse of the first step “wins” in
the competition for [A*]. However, in the other limit, that k ≫ k [A] , the reaction becomes second order in [A] since
2 −1
k−1 [A] + k ≈ k .
2 2
d[P ]
2
= k1 [A]
dt
which is consistent with the forward reaction of the first step being the rate determining step, since A
∗
is removed from the
reaction (through the formation of products) very quickly as soon as it is formed.
Third-body Collisions
Sometimes, the third-body collision is provided by an inert species M , perhaps by filling the reaction chamber with a heavy non-
reactive species, such as Ar. In this case, the mechanism becomes
k−1
∗
A+M ⇌ A +M
k1
k2
∗
A −
→P
And if the concentration of the third body collider is constant, it is convenient to define an effective rate constant, k uni .
12.7.1 https://chem.libretexts.org/@go/page/84601
k2 k1 [M ]
kuni =
k−1 [M ] + k2
The utility is that important information about the individual step rate constants can be extracted by plotting 1/k uni as a function of
1/[M ].
1 k−1 1
= + k2 ( )
kuni k2 k1 [M ]
The plot should yield a straight line, the slope of which gives the value of k , and the intercept gives (k
2 −1 / k2 k1 ).
This page titled 12.7: The Lindemann Mechanism is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
12.7.2 https://chem.libretexts.org/@go/page/84601
12.8: The Michaelis-Menten Mechanism
The Michaelis-Menten mechanism (Michaelis & Menten, 1913) is one which many enzyme mitigated reactions follow. The basic
mechanism involves an enzyme (E , a biological catalyst) and a substrate (S ) which must connect to form an enzyme-substrate
complex (ES ) in order for the substrate to be degraded (or augmented) to form a product (P ). The overall reaction is
S → P
E +S ⇌ ES (12.8.1)
k−1
k2
ES −
→P (12.8.2)
Notice that the enzyme is necessary for the reaction to proceed, but is not part of the overall stoichiometry (as is the case for any
catalyst!).
And using a mass conservation relationship on the enzyme (noting that the enzyme must be either in its bare form (E ) or
complexed with a substrate (ES )):
[E ]o = [E] + [ES]
or
[E] = [E ]o − [ES]
= (k−1 + k1 )[ES]
k1 [E ]o [S]
= [ES]
k1 [S] + k−1
Substituting this into the expression for the rate of production of the product P
d[P ]
= k2 [ES] (12.8.4)
dt
yields
d[P ] k2 k1 [E ]o [S]
=
dt k1 [S] + k−1
Multiplying the top and bottom of the expression on the right hand side by 1/k1 gives the result
d[P ] k2 [E ]o [S]
=
k1
dt
[S] +
k−1
12.8.1 https://chem.libretexts.org/@go/page/84602
The ratio of k-1/k1 is the equilibrium constant that describes the dissociation of the enzyme-substrate complex, Kd in Equation
d[P ]
12.8.1 . Noting that k2 [E ]0 gives the maximum rate (V max ), and that is the observed reaction rate, the rate law takes the form
dt
Vmax [S]
rate =
Kd + [S]
This is because the maximum reaction rate is achieved when [ES] is low. As [ES] increases, the likelihood of the complex
decomposing to reform reactants is higher, slowing the conversion. [ES] will be low if the concentration of the enzyme is
much larger than that of the substrate, so there is never a shortage of enzyme available to form the complex with the substrate.
However, if the substrate concentration is higher, the lack of available enzyme active sites will slow the reaction and cause it to
become 0th order.
In the limit that the substrate concentration is large compared to Kd (i.e., Kd + [S] ≈ [S] ), the reaction ends up zeroth order in
substrate.
Vmax [S]
Vmax [S]
rate = ≈ = Vmax
Kd + [S]
[S]
Hence, adding more substrate to the system under this limiting condition will have no effect on the observed rate. This is
characteristic of a bottleneck in the mechanism, which would happen if there is a shortage of enzyme sites to which the substrate
can attach.
In the other extreme, in which Kd is very large compared to the substrate concentration (i.e., Kd + [S] ≈ Kd ), the reaction
become first order in substrate.
Vmax [S] Vmax [S] Vmax
rate = ≈ = [S]
Kd + [S] Kd Kd
or
[E][S]
[ES] =
Km
where
k−1 + K2
Km =
k1
KM is the Michaelis constant, which is affected by a number of factors, including pH, temperature, and the nature of the substrate
itself. Proceeding as before, though the conservation of mass relationship and substitution into the expression for rate (Equation
12.8.4) results in
The advantage to this approach is that it accounts for the loss of ES complex due to the production of products as well as the
decomposition to reform the reactants E and S. As before, in the limit that [S] ≫ K , the reaction reaches its maximum rate (
M
12.8.2 https://chem.libretexts.org/@go/page/84602
Vmax ) and becomes independent of any concentrations. However in the limit that [S] ≪ K , the reaction becomes 1st order in [S].
M
The Michalis constant and Vmax parameters can be extracted in a number of ways. In the Lineweaver-Burk (Lineweaver & Burk,
1934) method, the reciprocal of the rate law is used to create a linear relationship.
1 Km + [S]
=
rate Vmax [S]
or
1 Km 1 1
= +
rate Vmax [S] Vmax
So a plot of 1/rate as a function of 1/[S] results in a straight line, the slope of which is equal to KM / Vmax and the intercept is
1/V max. This is called a Lineweaver–Burk plot.
This page titled 12.8: The Michaelis-Menten Mechanism is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated
by Patrick Fleming.
12.8.3 https://chem.libretexts.org/@go/page/84602
12.9: Chain Reactions
A large number of reactions proceed through a series of steps that can collectively be classified as a chain reaction. The reactions
contain steps that can be classified as
initiation step – a step that creates the intermediates from stable species
propagation step – a step that consumes an intermediate, but creates a new one
termination step – a step that consumes intermediates without creating new ones
These types of reactions are very common when the intermediates involved are radicals. An example, is the reaction
H2 + Br2 → 2H Br
A proposed mechanism is
k1
⋅
Br2 ⇌ 2Br (12.9.2)
k−1
k2
⋅ ⋅
2Br + H2 ⇌ H Br + H (12.9.3)
k−2
k3
⋅ ⋅
H + Br2 −
→ H Br + Br (12.9.4)
Based on this mechanism, the rate of change of concentrations for the intermediates (H and Br ) can be written, and the steady ⋅ ⋅
⋅
d[Br ]
⋅ 2 ⋅ ⋅ ⋅
= 2 k1 [Br2 ] − 2 k−1 [Br ] − k2 [Br ][ H2 ] + k−2 [H Br][ H ] + k3 [ H ][Br2 ] = 0
dt
Adding these two expressions cancels the terms involving k , k , and k . The result is
2 −2 3
⋅ 2
2 k1 [Br2 ] − 2 k−1 [Br ] =0
Solving for Br ⋅
−−−−−−−
k1 [Br2 ]
⋅
Br = √
k−1
This can be substituted into an expression for the H that is generated by solving the steady state expression for d[H
⋅ ⋅
]/dt .
⋅
k2 [Br ][ H2 ]
⋅
[H ] =
k−2 [H Br] + k3 [Br2 ]
so
−−−−−−−
k1 [Br2 ]
k2 √ [ H2 ]
k−1
⋅
[H ] =
k−2 [H Br] + k3 [Br2 ]
Now, armed with expressions for H and Br , we can substitute them into an expression for the rate of production of the product
⋅ ⋅
H Br :
[H Br]
⋅ ⋅ ⋅
= k2 [Br ][ H2 ] + k3 [ H ][Br2 ] − k−2 [ H ][H Br]
dt
12.9.1 https://chem.libretexts.org/@go/page/84603
After substitution and simplification, the result is
1/2
k1
1/2
2 k2 ( ) [ H2 ][Br2 ]
[H Br] k−1
=
dt k−1 [H Br]
1+
k3 [Br2 ]
Multiplying the top and bottom expressions on the right by [Br ] produces
2
1/2
k1
3/2
2 k2 ( ) [ H2 ][Br2 ]
[H Br] k−1
=
dt k−1
[Br2 ] + [H Br]
k3
which matches the form of the rate law found experimentally (Equation 12.9.1)! In this case,
−−−−
k1
k = 2k2 √
k−1
and
k−2
′
k =
k3
This page titled 12.9: Chain Reactions is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
12.9.2 https://chem.libretexts.org/@go/page/84603
12.10: Catalysis
There are many examples of reactions that involve catalysis. One that is of current importance to the chemistry of the environment
is the catalytic decomposition of ozone (Fahey, 2006). The overall reaction
⋅
O3 + O → 2 O2
O3 + C l −
→ C lO + O2
k1
C lO + O −
→ C l + O2
Applying the steady state approximation to this relationship and solving for [C lO] produces
[ O3 ][C l]
C lO] = (12.10.1)
k2 [O]
The rate of production of O (which is two times the rate of the reaction) is given by
2
d[ O2 ]
= k2 [ O3 ][C l] + k2 [C lO][O]
dt
Substituting the expression for [C lO] (Equation 12.10.1) into the above expression yields
d[ O2 ] [ O3 ][C l]
= k2 [ O3 ][C l] + k2 ( ) [O]
dt k2 [O]
= k1 [ O3 [C l] + k1 [ O3 ][C l]
= 2 k1 [ O3 ][C l]
And so the rate of the reaction is predicted to be first order in [O ], first order in the catalyst [C l], and second order overall.
3
rate = k[ O3 ][C l]
If the concentration of the catalyst is constant, the reaction kinetics will reduce to first order.
rate = k[ O3 ]
On the left, atomic oxygen picks up an oxygen atom from C lO to form O and generate a C l atom, which can then react with O
2 3
to form C lO and an O molecule. The closed loop in the middle is characteristic of the catalytic cycle involving C l and C lO.
2
Further, since C l acts as a catalyst, it can decompose many O molecules without being degraded through side reactions.
3
The introduction of chlorine atoms into the upper atmosphere is a major environmental problem, leading to the annual thinning and
eventual opening of the ozone layer over Antarctica. The source of chlorine is from the decomposition of chlorofluorocarbons
which are sued as refrigerants and propellants due to their incredible stability near the Earth’s surface. However, in the upper
12.10.1 https://chem.libretexts.org/@go/page/84604
atmosphere, these compounds are subjected to ultraviolet radiation emitted by the sun and decompose to form the radicals
responsible for the catalytic decomposition of ozone. The world community addressed this issue by drafting the Montreal Protocol
(Secretariat, 2015), which focused on the emission of ozone-destroying compounds. The result of this action has brought about
evidence of the Antarctic ozone hole healing (K, 2015). This is one very good example science-guided political, industrial, and
economic policies leading to positive changes for our environment.
This page titled 12.10: Catalysis is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick Fleming.
12.10.2 https://chem.libretexts.org/@go/page/84604
12.11: Oscillating Reactions
In most cases, the conversion of reactants into products is a fairly smooth process, in that the concentrations of the reactants
decrease in a regular manner, and those of the products increase in a similar regular manner. However, some reactions can show
irregular behavior in this regard. One particularly peculiar (but interesting!) phenomenon is that of oscillating reactions, in which
reactant concentrations can rise and fall as the reaction progresses. One way this can happen is when the products of the reaction
(or one of the steps) catalyzes the reaction (or one of the steps. This process is called autocatalysis.
An example of an autocatalyzed mechanism is the Lotka-Voltera mechanism. This is a three-step mechanism defined as follows:
k1
A+X −
→ X +X
k2
X +Y −
→ Y +Y
k3
Y −
→B
In this reaction, the concentration of reactant A is held constant by continually adding it to the reaction mixture. The first step is
autocatalyzed, so as it proceeds, it speeds up. However, an increase in the production of X by the first reaction increases the rate
of the second reaction as well, which is also autocatalyzed. Finally, the removal of Y through the third reaction brings things to a
halt, until the first reaction can again produce a build up of X to start the cycle over.
A plot of the concentration of X and Y as a function of time looks as follows:
This mechanism follows kinetics predicted by what is called the predator-prey relationship. In this case, X represents the “prey”
and Y represents the “predator”. The population of the predator cannot build up unless there is a significant population of prey on
which the predators can feed. Likewise, the population of predators decreases when the population of the prey falls. And finally,
there is a lag, as the rise and decline of the prey population controls the rise and fall of the predator population. The equations have
been studied extensively and have applications not just in chemical kinetics, but in biology, economics, and elsewhere. One
wonders if the equations can be applied to help to understand politics!
This page titled 12.11: Oscillating Reactions is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by Patrick
Fleming.
12.11.1 https://chem.libretexts.org/@go/page/84605
12.E: Chemical Kinetics II (Exercises)
In preparation
This page titled 12.E: Chemical Kinetics II (Exercises) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
12.E.1 https://chem.libretexts.org/@go/page/84608
12.S: Chemical Kinetics II (Summary)
Vocabulary and Concepts
autocatalysis
bimolecular
catalyst
chain reaction
dynamic equilibrium
effect rate constant
elementary reaction
equilibrium approximation
equilibrium constant
initiation step
intermediate
Lindemann mechanism
Lotka-Voltera mechanism
Michaelis constant
Michaelis-Menten
molecularity
Montreal Protocol
oscillating reaction
predator-prey
pre-equilibrium
propagation step
radical
rate determining step
reaction mechanism
steady state approximation
termination step
termolecular
third-body collision
unimolecular
References
1. Briggs, G., & Haldane, J. (1925). A note on the kinematics of enzyme action. Biochem J, 19(2), 338-339.
2. Fahey, D. W. (2006). Twenty Questions and Answers About the Ozone Layer: 2006 Update. Earth System Research Laboratory;
Chemical Sciences Division. Washington DC: U.S. Department of Commerce; National Oceanic and Atmospheric
Administration.
3. K, J. (2015, May 14). Ozone layer healing with 'hole' closing up says Nasa. Retrieved May 3, 2016, from International Business
Times: http://www.ibtimes.co.uk/ozone-layer...s-nasa-1501227
4. Lindemann, F. A., Arrhenius, S., Langmuir, I., Dhar, N. R., Perrin, J., & Lewis, W. C. (1922). Discussion on the "radiation
theory of chemical reaction". Trans. Faraday Soc., 17, 598-606.
5. Lineweaver, H., & Burk, D. (1934). The Determination of Enzyme Dissociation Constants. Journal of the American Chemical
Society, 56(3), 658-666.
6. Michaelis, L., & Menten, M. (1913). Die Kinetik der Invertinwirkung. Biochem Z, 49, 333-369.
7. Redmon, L. T., Purvis, G. D., & Bartlett, R. J. (1978). The unimolecular isomerization of methyl isocyanide to methyl cyanide
(acetonitrile). Hournal of Chemical Physics, 69, 5386.
8. Secretariat, O. (2015). THE MONTREAL PROTOCOL ON SUBSTANCES THAT DEPLETE THE OZONE LAYER . Retrieved
May 3, 2016, from The Ozone Secretariat: ozone.unep.org/en/treaties-an...te-ozone-layer
This page titled 12.S: Chemical Kinetics II (Summary) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
Patrick Fleming.
12.S.1 https://chem.libretexts.org/@go/page/84607
Index
A Charles' Law E
activated complex 2.1: The Empirical Gas Laws ebullioscopic constant
11.11: Transition State Theory Charles's Law 7.6: Colligative Properties
activation energy 2.1: The Empirical Gas Laws elementary reaction
11.9: Temperature Dependence chemical equilibria 12.1: Reaction Mechanisms
activity 9: Chemical Equilibria empirical gas laws
7.8: Non-ideality in Solutions - Activity chemical potential 2.1: The Empirical Gas Laws
9.3: Activities and Fugacities 7.3: Chemical Potential endothermic process
activity coefficient 9.2: Chemical Potential
3.2: Work and Heat
7.8: Non-ideality in Solutions - Activity Clapeyron Equation
energy
Adiabatic Changes 8.4: The Clapeyron Equation
1.5: Work and Energy
3.3: Reversible and Irreversible Pathways Clausius inequality
Enthalpy change
5.4: Calculating Entropy Changes 5.5: Comparing the System and the Surroundings
3.4: Calorimetry
adiabatic compressibility colligative property
enthalpy of combustion
5.8: Adiabatic Compressibility 7.6: Colligative Properties
3.4: Calorimetry
Arrhenius equation collision theory
enthalpy of mixing
11.9: Temperature Dependence 11.10: Collision Theory
7.1: Thermodynamics of Mixing
arrhenius prefactor collisional cross section
entropy
11.10: Collision Theory 2.6: Collisions with Other Molecules
11.10: Collision Theory 5.3: Entropy
autocatalysis entropy change of the surroundings
12.11: Oscillating Reactions Collisions with wall
2.4: Kinetic Energy 5.5: Comparing the System and the Surroundings
autoionization entropy change of the system
9.8: Acid-Base Equilibria combined gas law
2.1: The Empirical Gas Laws 5.5: Comparing the System and the Surroundings
average speed in gas Entropy of galvanic cell
2.3: The Kinetic Molecular Theory of Gases common ion effect
9.10: Solubility of Ionic Compounds 10.4: Entropy of Electrochemical Cells
Avogadro's law entropy of mixing
2.1: The Empirical Gas Laws compressibility factor
7.5: Non-ideality in Gases - Fugacity 7.1: Thermodynamics of Mixing
azeotrope EOS
8.8: Non-ideality - Henry's Law and Azeotropes compression factor
2.7: Real Gases 3.3: Reversible and Irreversible Pathways
concentration cell Euler relation
B 4.2: Total and Exact Differentials
10.5: Concentration Cells
bimolecular reaction eutectic halt
12.1: Reaction Mechanisms
Concentration Profiles
12.2: Concentration Profiles for Some Simple 8.10: Cooling Curves
boiling point elevation Mechanisms eutectic point
7.6: Colligative Properties
constant pressure heat capacity 8.9: Solid-Liquid Systems - Eutectic Points
Boltzmann equation for entropy 3.3: Reversible and Irreversible Pathways Exact Differentials
5.6: Entropy and Disorder 6.8: The Difference between Cp and Cv 4.2: Total and Exact Differentials
bomb calorimeter constant volume heat capacity exothermic
3.4: Calorimetry 3.3: Reversible and Irreversible Pathways 3.2: Work and Heat
bond energy 6.8: The Difference between Cp and Cv
EXPANDING THERMODYNAMIC
3.6: Reaction Enthalpies cooling curve
FUNCTIONS
Boyle temperature 8.2: Single Component Phase Diagrams
8.10: Cooling Curves 4.3: Compressibility and Expansivity
2.7: Real Gases 4.6: Useful Definitions and Relationships
Boyle's Law cryoscopic constant
7.6: Colligative Properties
extensive property
2.1: The Empirical Gas Laws 1.1: The System and the Surroundings
buffer cyclic permutation rule
4.3: Compressibility and Expansivity
9.9: Buffers
4.6: Useful Definitions and Relationships F
Faraday's constant
C D 10.1: Electricity
calorimetry first law of thermodynamics
Debye extrapolation
3.4: Calorimetry 3.2: Work and Heat
5.7: The Third Law of Thermodynamics
Carnot cycle first order reaction
degree of dissociation
5.2: Heat Engines and the Carnot Cycle 11.5: 1st order rate law
9.5: Degree of Dissociation
catalyst free energy
disorder
12.10: Catalysis 6.3: ΔA, ΔG, and Maximum Work
5.6: Entropy and Disorder
catalytic decomposition of ozone freezing point depression
distillation
12.10: Catalysis 7.6: Colligative Properties
8.7: Liquid-Vapor Systems - Raoult’s Law
cell potential freezing point depression constant
Dumas Bulb
10.3: Half Cells and Standard Reduction Potentials 7.6: Colligative Properties
9.7: The Dumas Bulb Method for Measuring
chain reaction Decomposition Equilibrium frequency of collisions
12.9: Chain Reactions 2.6: Collisions with Other Molecules
1 https://chem.libretexts.org/@go/page/212307
fugacity J O
7.5: Non-ideality in Gases - Fugacity Joule Experiment oscillating reactions
9.3: Activities and Fugacities
3.2: Work and Heat 12.11: Oscillating Reactions
fugacity coefficient 4.4: The Joule Experiment oxidation
7.5: Non-ideality in Gases - Fugacity
10.1: Electricity
K Ozone
G kelvin 12.10: Catalysis
galvanic cell 1.3: Temperature
10.3: Half Cells and Standard Reduction Potentials kinetic energy P
Gibbs energy of mixing 1.5: Work and Energy partial molar volume
7.1: Thermodynamics of Mixing 2.4: Kinetic Energy
7.2: Partial Molar Volume
Gibbs Function Kinetic Molecular Theory of Gases Partially Miscible Liquids
6.1: Free Energy Functions 2.3: The Kinetic Molecular Theory of Gases
8.6: Phase Diagrams for Binary Mixtures
Graham's law Kirchhoff's Law partition function
2.5: Graham’s Law of Effusion 3.5: Temperature Dependence of Enthalpy
11.11: Transition State Theory
Knudsen cell phase diagram
H 2.5: Graham’s Law of Effusion
8.1: Prelude to Phase Equilibrium
heat 8.2: Single Component Phase Diagrams
3.2: Work and Heat L Phase Rule
Heat capacity Lattice Energy 8.3: Criterion for Phase Equilibrium
5.2: Heat Engines and the Carnot Cycle 3.7: Lattice Energy and the Born-Haber Cycle Potential Energy
heat is not a state function Le Chatelier's Principle 1.5: Work and Energy
5.3: Entropy 9.4: Pressure Dependence of Kp - Le Châtelier's pressure
Helmholtz function Principle 1.2: Pressure and Molar Volume
6.1: Free Energy Functions lever rule Pressure Dependence of Equilibrium
Henry's law 8.6: Phase Diagrams for Binary Mixtures
Constant
8.8: Non-ideality - Henry's Law and Azeotropes Lindemann mechanism
9.4: Pressure Dependence of Kp - Le Châtelier's
Hess's law 12.7: The Lindemann Mechanism Principle
3.6: Reaction Enthalpies lower inversion temperature Pressure Dependence of Gibbs Energy
4.5: The Joule-Thomson Effect 6.5: Pressure Dependence of Gibbs Energy
I principle of corresponding states
ICE Table M 2.7: Real Gases
9.5: Degree of Dissociation maximum boiling point azeotrope propagation step
ideal gas law 8.8: Non-ideality - Henry's Law and Azeotropes 12.9: Chain Reactions
2.2: The Ideal Gas Law Maxwell Relations
Incongruent Melting 6.2: Combining the First and Second Laws - R
Maxwell's Relations
8.9: Solid-Liquid Systems - Eutectic Points Raoult's Law
inexact differentials mean activity coefficient
8.7: Liquid-Vapor Systems - Raoult’s Law
7.8: Non-ideality in Solutions - Activity
4.2: Total and Exact Differentials Rapid Equilibrium Approximation
initiation step mean free path
12.2: Concentration Profiles for Some Simple
2.6: Collisions with Other Molecules Mechanisms
12.9: Chain Reactions
intensive property Method of Initial Rates 12.3: The Connection between Reaction
11.7: The Method of Initial Rates Mechanisms and Reaction Rate Laws
1.1: The System and the Surroundings 12.6: The Equilibrium Approximation
internal energy Michaelis constant
12.8: The Michaelis-Menten Mechanism
rate constant
3.2: Work and Heat 11.1: Reaction Rate
internal pressure minimum boiling point azeotrope
8.8: Non-ideality - Henry's Law and Azeotropes
rate determining step
4.4: The Joule Experiment 12.3: The Connection between Reaction
ionic strength molecularity Mechanisms and Reaction Rate Laws
12.1: Reaction Mechanisms 12.4: The Rate Determining Step Approximation
7.8: Non-ideality in Solutions - Activity
ionization energy Montreal Protocol rate law
12.10: Catalysis 11.3: Rate Laws
3.6: Reaction Enthalpies
Isobaric Changes most probable speed in gas reaction enthalpy
2.3: The Kinetic Molecular Theory of Gases 3.6: Reaction Enthalpies
5.4: Calculating Entropy Changes
isobaric thermal expansivity reaction mechanism
4.3: Compressibility and Expansivity
N 12.1: Reaction Mechanisms
isochoric natural variable Reaction order
3.2: Work and Heat 6.2: Combining the First and Second Laws - 11.3: Rate Laws
Maxwell's Relations Reaction Quotient
isothermal
3.2: Work and Heat
Nernst Equation 9.2: Chemical Potential
10.2: The connection to ΔG reaction rate
Isothermal Changes in Entropy
5.4: Calculating Entropy Changes 11.1: Reaction Rate
isothermal compressibility reciprocal rule
4.3: Compressibility and Expansivity 4.3: Compressibility and Expansivity
2 https://chem.libretexts.org/@go/page/212307
reduction stopped flow U
10.1: Electricity 11.2: Measuring Reaction Rates unimolecular reaction
reversible expansion surroundings 12.1: Reaction Mechanisms
1.5: Work and Energy 1.1: The System and the Surroundings upper inversion temperature
RMS speed in gas system 4.5: The Joule-Thomson Effect
2.3: The Kinetic Molecular Theory of Gases 1.1: The System and the Surroundings
V
S T van’t Hoff equation
Second Law of Thermodynamics temperature 9.6: Temperature Dependence of Equilibrium
5.1: Introduction to the Second Law 1.3: Temperature Constants - the van ’t Hoff Equation
second order reaction Temperature Dependence of Equilibrium van’t Hoff plot
11.6: 2nd order Rate Laws Constant 9.6: Temperature Dependence of Equilibrium
Second Virial Coefficient 9.6: Temperature Dependence of Equilibrium
Constants - the van ’t Hoff Equation
2.7: Real Gases Constants - the van ’t Hoff Equation vapor pressure
solubility Temperature Dependence of Gibbs 7.6: Colligative Properties
7.7: Solubility Energy viral EOS
solubility product 6.6: Temperature Dependence of A and G
2.7: Real Gases
9.10: Solubility of Ionic Compounds Temperature Dependence of Helmholtz Volume Dependence of Helmholtz
specific heat Energy Energy
1.5: Work and Energy 6.4: Volume Dependence of Helmholtz Energy
6.6: Temperature Dependence of A and G
spectrophotometry Temperature Dependence to Rates
11.2: Measuring Reaction Rates
11.9: Temperature Dependence
W
speed of sound termination step weak acid
5.8: Adiabatic Compressibility 9.8: Acid-Base Equilibria
12.9: Chain Reactions
spontaneous change termolecular reaction weak bases
5.1: Introduction to the Second Law 9.8: Acid-Base Equilibria
12.1: Reaction Mechanisms
standard cell potential third law entropy work
10.2: The connection to ΔG 1.5: Work and Energy
5.7: The Third Law of Thermodynamics
Standard Enthalpy of Formation Third Law of Thermodynamics 3.2: Work and Heat
3.6: Reaction Enthalpies
5.7: The Third Law of Thermodynamics
standard reduction potential Torricelli barometer Z
10.3: Half Cells and Standard Reduction Potentials
1.2: Pressure and Molar Volume Zero Law of Thermodynamics
standard state Total Differentials 1.4: The Zeroth Law of Thermodynamics
3.6: Reaction Enthalpies
4.2: Total and Exact Differentials zeroth order reaction
state function transition state 11.4: 0th order Rate Law
1.1: The System and the Surroundings
11.11: Transition State Theory
steady state approximation triple point
12.2: Concentration Profiles for Some Simple
8.3: Criterion for Phase Equilibrium
Mechanisms
12.3: The Connection between Reaction Two Variables Change at Once
Mechanisms and Reaction Rate Laws 6.7: When Two Variables Change at Once
12.5: The Steady-State Approximation
3 https://chem.libretexts.org/@go/page/212307
Index
A Charles' Law E
activated complex 2.1: The Empirical Gas Laws ebullioscopic constant
11.11: Transition State Theory Charles's Law 7.6: Colligative Properties
activation energy 2.1: The Empirical Gas Laws elementary reaction
11.9: Temperature Dependence chemical equilibria 12.1: Reaction Mechanisms
activity 9: Chemical Equilibria empirical gas laws
7.8: Non-ideality in Solutions - Activity chemical potential 2.1: The Empirical Gas Laws
9.3: Activities and Fugacities 7.3: Chemical Potential endothermic process
activity coefficient 9.2: Chemical Potential
3.2: Work and Heat
7.8: Non-ideality in Solutions - Activity Clapeyron Equation
energy
Adiabatic Changes 8.4: The Clapeyron Equation
1.5: Work and Energy
3.3: Reversible and Irreversible Pathways Clausius inequality
Enthalpy change
5.4: Calculating Entropy Changes 5.5: Comparing the System and the Surroundings
3.4: Calorimetry
adiabatic compressibility colligative property
enthalpy of combustion
5.8: Adiabatic Compressibility 7.6: Colligative Properties
3.4: Calorimetry
Arrhenius equation collision theory
enthalpy of mixing
11.9: Temperature Dependence 11.10: Collision Theory
7.1: Thermodynamics of Mixing
arrhenius prefactor collisional cross section
entropy
11.10: Collision Theory 2.6: Collisions with Other Molecules
11.10: Collision Theory 5.3: Entropy
autocatalysis entropy change of the surroundings
12.11: Oscillating Reactions Collisions with wall
2.4: Kinetic Energy 5.5: Comparing the System and the Surroundings
autoionization entropy change of the system
9.8: Acid-Base Equilibria combined gas law
2.1: The Empirical Gas Laws 5.5: Comparing the System and the Surroundings
average speed in gas Entropy of galvanic cell
2.3: The Kinetic Molecular Theory of Gases common ion effect
9.10: Solubility of Ionic Compounds 10.4: Entropy of Electrochemical Cells
Avogadro's law entropy of mixing
2.1: The Empirical Gas Laws compressibility factor
7.5: Non-ideality in Gases - Fugacity 7.1: Thermodynamics of Mixing
azeotrope EOS
8.8: Non-ideality - Henry's Law and Azeotropes compression factor
2.7: Real Gases 3.3: Reversible and Irreversible Pathways
concentration cell Euler relation
B 4.2: Total and Exact Differentials
10.5: Concentration Cells
bimolecular reaction eutectic halt
12.1: Reaction Mechanisms
Concentration Profiles
12.2: Concentration Profiles for Some Simple 8.10: Cooling Curves
boiling point elevation Mechanisms eutectic point
7.6: Colligative Properties
constant pressure heat capacity 8.9: Solid-Liquid Systems - Eutectic Points
Boltzmann equation for entropy 3.3: Reversible and Irreversible Pathways Exact Differentials
5.6: Entropy and Disorder 6.8: The Difference between Cp and Cv 4.2: Total and Exact Differentials
bomb calorimeter constant volume heat capacity exothermic
3.4: Calorimetry 3.3: Reversible and Irreversible Pathways 3.2: Work and Heat
bond energy 6.8: The Difference between Cp and Cv
EXPANDING THERMODYNAMIC
3.6: Reaction Enthalpies cooling curve
FUNCTIONS
Boyle temperature 8.2: Single Component Phase Diagrams
8.10: Cooling Curves 4.3: Compressibility and Expansivity
2.7: Real Gases 4.6: Useful Definitions and Relationships
Boyle's Law cryoscopic constant
7.6: Colligative Properties
extensive property
2.1: The Empirical Gas Laws 1.1: The System and the Surroundings
buffer cyclic permutation rule
4.3: Compressibility and Expansivity
9.9: Buffers
4.6: Useful Definitions and Relationships F
Faraday's constant
C D 10.1: Electricity
calorimetry first law of thermodynamics
Debye extrapolation
3.4: Calorimetry 3.2: Work and Heat
5.7: The Third Law of Thermodynamics
Carnot cycle first order reaction
degree of dissociation
5.2: Heat Engines and the Carnot Cycle 11.5: 1st order rate law
9.5: Degree of Dissociation
catalyst free energy
disorder
12.10: Catalysis 6.3: ΔA, ΔG, and Maximum Work
5.6: Entropy and Disorder
catalytic decomposition of ozone freezing point depression
distillation
12.10: Catalysis 7.6: Colligative Properties
8.7: Liquid-Vapor Systems - Raoult’s Law
cell potential freezing point depression constant
Dumas Bulb
10.3: Half Cells and Standard Reduction Potentials 7.6: Colligative Properties
9.7: The Dumas Bulb Method for Measuring
chain reaction Decomposition Equilibrium frequency of collisions
12.9: Chain Reactions 2.6: Collisions with Other Molecules
1 https://chem.libretexts.org/@go/page/350621
fugacity J O
7.5: Non-ideality in Gases - Fugacity Joule Experiment oscillating reactions
9.3: Activities and Fugacities
3.2: Work and Heat 12.11: Oscillating Reactions
fugacity coefficient 4.4: The Joule Experiment oxidation
7.5: Non-ideality in Gases - Fugacity
10.1: Electricity
K Ozone
G kelvin 12.10: Catalysis
galvanic cell 1.3: Temperature
10.3: Half Cells and Standard Reduction Potentials kinetic energy P
Gibbs energy of mixing 1.5: Work and Energy partial molar volume
7.1: Thermodynamics of Mixing 2.4: Kinetic Energy
7.2: Partial Molar Volume
Gibbs Function Kinetic Molecular Theory of Gases Partially Miscible Liquids
6.1: Free Energy Functions 2.3: The Kinetic Molecular Theory of Gases
8.6: Phase Diagrams for Binary Mixtures
Graham's law Kirchhoff's Law partition function
2.5: Graham’s Law of Effusion 3.5: Temperature Dependence of Enthalpy
11.11: Transition State Theory
Knudsen cell phase diagram
H 2.5: Graham’s Law of Effusion
8.1: Prelude to Phase Equilibrium
heat 8.2: Single Component Phase Diagrams
3.2: Work and Heat L Phase Rule
Heat capacity Lattice Energy 8.3: Criterion for Phase Equilibrium
5.2: Heat Engines and the Carnot Cycle 3.7: Lattice Energy and the Born-Haber Cycle Potential Energy
heat is not a state function Le Chatelier's Principle 1.5: Work and Energy
5.3: Entropy 9.4: Pressure Dependence of Kp - Le Châtelier's pressure
Helmholtz function Principle 1.2: Pressure and Molar Volume
6.1: Free Energy Functions lever rule Pressure Dependence of Equilibrium
Henry's law 8.6: Phase Diagrams for Binary Mixtures
Constant
8.8: Non-ideality - Henry's Law and Azeotropes Lindemann mechanism
9.4: Pressure Dependence of Kp - Le Châtelier's
Hess's law 12.7: The Lindemann Mechanism Principle
3.6: Reaction Enthalpies lower inversion temperature Pressure Dependence of Gibbs Energy
4.5: The Joule-Thomson Effect 6.5: Pressure Dependence of Gibbs Energy
I principle of corresponding states
ICE Table M 2.7: Real Gases
9.5: Degree of Dissociation maximum boiling point azeotrope propagation step
ideal gas law 8.8: Non-ideality - Henry's Law and Azeotropes 12.9: Chain Reactions
2.2: The Ideal Gas Law Maxwell Relations
Incongruent Melting 6.2: Combining the First and Second Laws - R
Maxwell's Relations
8.9: Solid-Liquid Systems - Eutectic Points Raoult's Law
inexact differentials mean activity coefficient
8.7: Liquid-Vapor Systems - Raoult’s Law
7.8: Non-ideality in Solutions - Activity
4.2: Total and Exact Differentials Rapid Equilibrium Approximation
initiation step mean free path
12.2: Concentration Profiles for Some Simple
2.6: Collisions with Other Molecules Mechanisms
12.9: Chain Reactions
intensive property Method of Initial Rates 12.3: The Connection between Reaction
11.7: The Method of Initial Rates Mechanisms and Reaction Rate Laws
1.1: The System and the Surroundings 12.6: The Equilibrium Approximation
internal energy Michaelis constant
12.8: The Michaelis-Menten Mechanism
rate constant
3.2: Work and Heat 11.1: Reaction Rate
internal pressure minimum boiling point azeotrope
8.8: Non-ideality - Henry's Law and Azeotropes
rate determining step
4.4: The Joule Experiment 12.3: The Connection between Reaction
ionic strength molecularity Mechanisms and Reaction Rate Laws
12.1: Reaction Mechanisms 12.4: The Rate Determining Step Approximation
7.8: Non-ideality in Solutions - Activity
ionization energy Montreal Protocol rate law
12.10: Catalysis 11.3: Rate Laws
3.6: Reaction Enthalpies
Isobaric Changes most probable speed in gas reaction enthalpy
2.3: The Kinetic Molecular Theory of Gases 3.6: Reaction Enthalpies
5.4: Calculating Entropy Changes
isobaric thermal expansivity reaction mechanism
4.3: Compressibility and Expansivity
N 12.1: Reaction Mechanisms
isochoric natural variable Reaction order
3.2: Work and Heat 6.2: Combining the First and Second Laws - 11.3: Rate Laws
Maxwell's Relations Reaction Quotient
isothermal
3.2: Work and Heat
Nernst Equation 9.2: Chemical Potential
10.2: The connection to ΔG reaction rate
Isothermal Changes in Entropy
5.4: Calculating Entropy Changes 11.1: Reaction Rate
isothermal compressibility reciprocal rule
4.3: Compressibility and Expansivity 4.3: Compressibility and Expansivity
2 https://chem.libretexts.org/@go/page/350621
reduction stopped flow U
10.1: Electricity 11.2: Measuring Reaction Rates unimolecular reaction
reversible expansion surroundings 12.1: Reaction Mechanisms
1.5: Work and Energy 1.1: The System and the Surroundings upper inversion temperature
RMS speed in gas system 4.5: The Joule-Thomson Effect
2.3: The Kinetic Molecular Theory of Gases 1.1: The System and the Surroundings
V
S T van’t Hoff equation
Second Law of Thermodynamics temperature 9.6: Temperature Dependence of Equilibrium
5.1: Introduction to the Second Law 1.3: Temperature Constants - the van ’t Hoff Equation
second order reaction Temperature Dependence of Equilibrium van’t Hoff plot
11.6: 2nd order Rate Laws Constant 9.6: Temperature Dependence of Equilibrium
Second Virial Coefficient 9.6: Temperature Dependence of Equilibrium
Constants - the van ’t Hoff Equation
2.7: Real Gases Constants - the van ’t Hoff Equation vapor pressure
solubility Temperature Dependence of Gibbs 7.6: Colligative Properties
7.7: Solubility Energy viral EOS
solubility product 6.6: Temperature Dependence of A and G
2.7: Real Gases
9.10: Solubility of Ionic Compounds Temperature Dependence of Helmholtz Volume Dependence of Helmholtz
specific heat Energy Energy
1.5: Work and Energy 6.4: Volume Dependence of Helmholtz Energy
6.6: Temperature Dependence of A and G
spectrophotometry Temperature Dependence to Rates
11.2: Measuring Reaction Rates
11.9: Temperature Dependence
W
speed of sound termination step weak acid
5.8: Adiabatic Compressibility 9.8: Acid-Base Equilibria
12.9: Chain Reactions
spontaneous change termolecular reaction weak bases
5.1: Introduction to the Second Law 9.8: Acid-Base Equilibria
12.1: Reaction Mechanisms
standard cell potential third law entropy work
10.2: The connection to ΔG 1.5: Work and Energy
5.7: The Third Law of Thermodynamics
Standard Enthalpy of Formation Third Law of Thermodynamics 3.2: Work and Heat
3.6: Reaction Enthalpies
5.7: The Third Law of Thermodynamics
standard reduction potential Torricelli barometer Z
10.3: Half Cells and Standard Reduction Potentials
1.2: Pressure and Molar Volume Zero Law of Thermodynamics
standard state Total Differentials 1.4: The Zeroth Law of Thermodynamics
3.6: Reaction Enthalpies
4.2: Total and Exact Differentials zeroth order reaction
state function transition state 11.4: 0th order Rate Law
1.1: The System and the Surroundings
11.11: Transition State Theory
steady state approximation triple point
12.2: Concentration Profiles for Some Simple
8.3: Criterion for Phase Equilibrium
Mechanisms
12.3: The Connection between Reaction Two Variables Change at Once
Mechanisms and Reaction Rate Laws 6.7: When Two Variables Change at Once
12.5: The Steady-State Approximation
3 https://chem.libretexts.org/@go/page/350621
Glossary
Sample Word 1 | Sample Definition 1
1 https://chem.libretexts.org/@go/page/279584
Detailed Licensing
Overview
Title: Physical Chemistry (Fleming)
Webpages: 143
Applicable Restrictions: Noncommercial
All licenses found:
CC BY-NC-SA 4.0: 97.9% (140 pages)
Undeclared: 2.1% (3 pages)
By Page
Physical Chemistry (Fleming) - CC BY-NC-SA 4.0 3.7: Lattice Energy and the Born-Haber Cycle - CC
Front Matter - CC BY-NC-SA 4.0 BY-NC-SA 4.0
TitlePage - CC BY-NC-SA 4.0 3.E: First Law of Thermodynamics (Exercises) - CC
InfoPage - CC BY-NC-SA 4.0 BY-NC-SA 4.0
Table of Contents - Undeclared 3.S: First Law of Thermodynamics (Summary) - CC
Licensing - Undeclared BY-NC-SA 4.0
4: Putting the First Law to Work - CC BY-NC-SA 4.0
1: The Basics - CC BY-NC-SA 4.0
4.1: Prelude to Putting the First Law to Work - CC
1.1: The System and the Surroundings - CC BY-NC-
BY-NC-SA 4.0
SA 4.0
4.2: Total and Exact Differentials - CC BY-NC-SA 4.0
1.2: Pressure and Molar Volume - CC BY-NC-SA 4.0
4.3: Compressibility and Expansivity - CC BY-NC-SA
1.3: Temperature - CC BY-NC-SA 4.0
4.0
1.4: The Zeroth Law of Thermodynamics - CC BY-
4.4: The Joule Experiment - CC BY-NC-SA 4.0
NC-SA 4.0
4.5: The Joule-Thomson Effect - CC BY-NC-SA 4.0
1.5: Work and Energy - CC BY-NC-SA 4.0
4.6: Useful Definitions and Relationships - CC BY-
1.E: The Basics (Exercises) - CC BY-NC-SA 4.0
NC-SA 4.0
1.S: The Basics (Summary) - CC BY-NC-SA 4.0
4.E: Putting the First Law to Work (Exercises) - CC
2: Gases - CC BY-NC-SA 4.0
BY-NC-SA 4.0
2.1: The Empirical Gas Laws - CC BY-NC-SA 4.0 4.S: Putting the First Law to Work (Summary) - CC
2.2: The Ideal Gas Law - CC BY-NC-SA 4.0 BY-NC-SA 4.0
2.3: The Kinetic Molecular Theory of Gases - CC BY-
5: The Second Law - CC BY-NC-SA 4.0
NC-SA 4.0
2.4: Kinetic Energy - CC BY-NC-SA 4.0 5.1: Introduction to the Second Law - CC BY-NC-SA
2.5: Graham’s Law of Effusion - CC BY-NC-SA 4.0 4.0
2.6: Collisions with Other Molecules - CC BY-NC-SA 5.2: Heat Engines and the Carnot Cycle - CC BY-NC-
4.0 SA 4.0
2.7: Real Gases - CC BY-NC-SA 4.0 5.3: Entropy - CC BY-NC-SA 4.0
2.E: Gases (Exercises) - CC BY-NC-SA 4.0 5.4: Calculating Entropy Changes - CC BY-NC-SA 4.0
2.S: Gases (Summary) - CC BY-NC-SA 4.0 5.5: Comparing the System and the Surroundings -
CC BY-NC-SA 4.0
3: First Law of Thermodynamics - CC BY-NC-SA 4.0
5.6: Entropy and Disorder - CC BY-NC-SA 4.0
3.1: Prelude to Thermodynamics - CC BY-NC-SA 4.0 5.7: The Third Law of Thermodynamics - CC BY-NC-
3.2: Work and Heat - CC BY-NC-SA 4.0 SA 4.0
3.3: Reversible and Irreversible Pathways - CC BY- 5.8: Adiabatic Compressibility - CC BY-NC-SA 4.0
NC-SA 4.0 5.E: The Second Law (Exercises) - CC BY-NC-SA 4.0
3.4: Calorimetry - CC BY-NC-SA 4.0 5.S: The Second Law (Summary) - CC BY-NC-SA 4.0
3.5: Temperature Dependence of Enthalpy - CC BY-
6: Putting the Second Law to Work - CC BY-NC-SA 4.0
NC-SA 4.0
3.6: Reaction Enthalpies - CC BY-NC-SA 4.0 6.1: Free Energy Functions - CC BY-NC-SA 4.0
1 https://chem.libretexts.org/@go/page/417262
6.2: Combining the First and Second Laws - 8.E: Phase Equilibrium (Exercises) - CC BY-NC-SA
Maxwell's Relations - CC BY-NC-SA 4.0 4.0
6.3: ΔA, ΔG, and Maximum Work - CC BY-NC-SA 8.S: Phase Equilibrium (Summary) - CC BY-NC-SA
4.0 4.0
6.4: Volume Dependence of Helmholtz Energy - CC 9: Chemical Equilibria - CC BY-NC-SA 4.0
BY-NC-SA 4.0 9.1: Prelude to Chemical Equilibria - CC BY-NC-SA
6.5: Pressure Dependence of Gibbs Energy - CC BY- 4.0
NC-SA 4.0 9.2: Chemical Potential - CC BY-NC-SA 4.0
6.6: Temperature Dependence of A and G - CC BY- 9.3: Activities and Fugacities - CC BY-NC-SA 4.0
NC-SA 4.0 9.4: Pressure Dependence of Kp - Le Châtelier's
6.7: When Two Variables Change at Once - CC BY- Principle - CC BY-NC-SA 4.0
NC-SA 4.0 9.5: Degree of Dissociation - CC BY-NC-SA 4.0
6.8: The Difference between Cp and Cv - CC BY-NC- 9.6: Temperature Dependence of Equilibrium
SA 4.0 Constants - the van ’t Hoff Equation - CC BY-NC-SA
6.E: Putting the Second Law to Work (Exercises) - 4.0
CC BY-NC-SA 4.0 9.7: The Dumas Bulb Method for Measuring
6.S: Putting the Second Law to Work (Summary) - Decomposition Equilibrium - CC BY-NC-SA 4.0
CC BY-NC-SA 4.0 9.8: Acid-Base Equilibria - CC BY-NC-SA 4.0
7: Mixtures and Solutions - CC BY-NC-SA 4.0 9.9: Buffers - CC BY-NC-SA 4.0
7.1: Thermodynamics of Mixing - CC BY-NC-SA 4.0 9.10: Solubility of Ionic Compounds - CC BY-NC-SA
7.2: Partial Molar Volume - CC BY-NC-SA 4.0 4.0
7.3: Chemical Potential - CC BY-NC-SA 4.0 9.E: Chemical Equilibria (Exercises) - CC BY-NC-SA
7.4: The Gibbs-Duhem Equation - CC BY-NC-SA 4.0 4.0
7.5: Non-ideality in Gases - Fugacity - CC BY-NC-SA 9.S: Chemical Equilibria (Summary) - CC BY-NC-SA
4.0 4.0
7.6: Colligative Properties - CC BY-NC-SA 4.0 10: Electrochemistry - CC BY-NC-SA 4.0
7.7: Solubility - CC BY-NC-SA 4.0
10.1: Electricity - CC BY-NC-SA 4.0
7.8: Non-ideality in Solutions - Activity - CC BY-NC-
10.2: The connection to ΔG - CC BY-NC-SA 4.0
SA 4.0
10.3: Half Cells and Standard Reduction Potentials -
7.E: Mixtures and Solutions (Exercises) - CC BY-NC-
CC BY-NC-SA 4.0
SA 4.0
10.4: Entropy of Electrochemical Cells - CC BY-NC-
7.S: Mixtures and Solutions (Summary) - CC BY-NC-
SA 4.0
SA 4.0
10.5: Concentration Cells - CC BY-NC-SA 4.0
8: Phase Equilibrium - CC BY-NC-SA 4.0 10.E: Electrochemistry (Exercises) - CC BY-NC-SA
8.1: Prelude to Phase Equilibrium - CC BY-NC-SA 4.0 4.0
8.2: Single Component Phase Diagrams - CC BY-NC- 10.S: Electrochemistry (Summary) - CC BY-NC-SA
SA 4.0 4.0
8.3: Criterion for Phase Equilibrium - CC BY-NC-SA 11: Chemical Kinetics I - CC BY-NC-SA 4.0
4.0 11.1: Reaction Rate - CC BY-NC-SA 4.0
8.4: The Clapeyron Equation - CC BY-NC-SA 4.0 11.2: Measuring Reaction Rates - CC BY-NC-SA 4.0
8.5: The Clausius-Clapeyron Equation - CC BY-NC- 11.3: Rate Laws - CC BY-NC-SA 4.0
SA 4.0 11.4: 0th order Rate Law - CC BY-NC-SA 4.0
8.6: Phase Diagrams for Binary Mixtures - CC BY- 11.5: 1st order rate law - CC BY-NC-SA 4.0
NC-SA 4.0 11.6: 2nd order Rate Laws - CC BY-NC-SA 4.0
8.7: Liquid-Vapor Systems - Raoult’s Law - CC BY- 11.7: The Method of Initial Rates - CC BY-NC-SA 4.0
NC-SA 4.0 11.8: The Method of Half-Lives - CC BY-NC-SA 4.0
8.8: Non-ideality - Henry's Law and Azeotropes - CC 11.9: Temperature Dependence - CC BY-NC-SA 4.0
BY-NC-SA 4.0 11.10: Collision Theory - CC BY-NC-SA 4.0
8.9: Solid-Liquid Systems - Eutectic Points - CC BY- 11.11: Transition State Theory - CC BY-NC-SA 4.0
NC-SA 4.0 11.E: Chemical Kinetics I (Exercises) - CC BY-NC-
8.10: Cooling Curves - CC BY-NC-SA 4.0 SA 4.0
2 https://chem.libretexts.org/@go/page/417262
11.S: Chemical Kinetics I (Summary) - CC BY-NC- 12.8: The Michaelis-Menten Mechanism - CC BY-
SA 4.0 NC-SA 4.0
12: Chemical Kinetics II - CC BY-NC-SA 4.0 12.9: Chain Reactions - CC BY-NC-SA 4.0
12.1: Reaction Mechanisms - CC BY-NC-SA 4.0 12.10: Catalysis - CC BY-NC-SA 4.0
12.2: Concentration Profiles for Some Simple 12.11: Oscillating Reactions - CC BY-NC-SA 4.0
Mechanisms - CC BY-NC-SA 4.0 12.E: Chemical Kinetics II (Exercises) - CC BY-NC-
12.3: The Connection between Reaction Mechanisms SA 4.0
and Reaction Rate Laws - CC BY-NC-SA 4.0 12.S: Chemical Kinetics II (Summary) - CC BY-NC-
12.4: The Rate Determining Step Approximation - SA 4.0
CC BY-NC-SA 4.0 Back Matter - CC BY-NC-SA 4.0
12.5: The Steady-State Approximation - CC BY-NC- Index - CC BY-NC-SA 4.0
SA 4.0 Index - CC BY-NC-SA 4.0
12.6: The Equilibrium Approximation - CC BY-NC- Glossary - CC BY-NC-SA 4.0
SA 4.0 Detailed Licensing - Undeclared
12.7: The Lindemann Mechanism - CC BY-NC-SA 4.0
3 https://chem.libretexts.org/@go/page/417262