0% found this document useful (0 votes)
817 views311 pages

Full PDF

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
817 views311 pages

Full PDF

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 311

UC DAVIS: CHEM 2C

GENERAL CHEMISTRY
III

LibreTexts
UC Davis: Chem 2C General Chemistry iiI

LibreTexts
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 MindTouch® 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. Unless otherwise noted, LibreTexts content is licensed by CC BY-NC-SA 3.0.
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 03/12/2023


TABLE OF CONTENTS

AGENDA
Grading on a Curve

Text
Licensing
1: Chemistry Primer
1.1: Oxidation-Reduction Reactions
1.2: Balancing Redox Reactions
1.3: Free Energy and Equilibrium (quick)
1.4: Free Energy and Equilibrium (complete)
1.5: Electronic Con gurations - Spin Quantum Number
1.6: Electronic Con gurations - Hund's Rule
1.7: Electronic Con gurations - The Aufbau Process
1.8: Electronic Con gurations - Pauli Exclusion Principle
2: Electrochemistry
2.1: Galvanic Cells
2.2: Standard Electrode Potentials
2.3: Strengths of Oxidants and Reductants
2.4: E, ΔG, and K (The Big Triangle of Chemistry)
2.5: Cell Potential as a Function of Concentrations
2.6: Batteries- Producing Electricity Through Chemical Reactions
2.7: Corrosion- Unwanted Voltaic Cells
2.8: Electrolysis- Causing Nonspontaneous Reactions to Occur
3: Coordination Chemistry
3.1: Werner’s Theory of Coordination Compounds
3.2: Ligands
3.3: Nomenclature
3.4: Isomerism
3.5: Optical Isomerism in Metal Complexes
3.6: Stability Aspects of Complex-Ion Equilibria
3.7: Chelation
3.8: Bonding in Octahedral Complex Ions- Crystal Field Theory
3.9: Bonding in Non-Octahedral Complex Ions- Crystal Field Theory
3.10: Magnetic Behavior of Atoms, Molecules, and Materials
3.11: Magnetic Behavior of Complex Ions
3.12: Optical Properties of Coordination Compounds (Color)
4: Descriptive Chemistry
5: Chemical Kinetics
5.1: Factors that Affect Reaction Rates
5.2: Reaction Rates
5.3: Concentration and Rates (Rate Laws)
5.4: Determining Rate Laws from Initial Rates (Differential Rate Laws)
5.5: The Change of Concentration with Time (Integrated Rate Laws)
5.6: Using Graphs to Determine (Integrated) Rate Laws
5.7: Collision Theory

1 https://chem.libretexts.org/@go/page/182232
5.8: Temperature and Rate
5.9: Reaction Mechanisms
5.10: Catalysis
5.11: Exercises
5.12: Chemical Kinetics (Summary)
6: Nuclear Chemistry
6.1: Components of the Nucleus
6.2: Nuclear Reactions
6.3: Nuclear Radiation
6.4: Rates of Radioactive Decay
6.5: Stability of the Atomic Nucleus
6.6: The Origin of the Elements
6.7: Transmutation of the Elements
6.8: Nuclear Fission
6.9: Nuclear Fusion
6.10: Nuclear Chemistry (Exercises)

Worksheets
1: Balancing Redox Reactions (Worksheet)
2: Galvanic Cells (Worksheet)
3: Complex Ions and Nomenclature (Worksheet)
4: Complex Ion Structure and Geometry (Worksheet)
5: Crystal Field Theory (Worksheet)
6: Magnetism and Colors in Coordination Complexes (Worksheet)
7: Kinetics I (Worksheet)
8: Kinetics II (Worksheet)
9: Kinetics III (Worksheet)
10: Fundamentals of Nuclear Chemistry (WorkSheet)
11: Transmutation and Nuclear Kinetics (Worksheet)

Student Academic Success Center Workshops


Additional Resources: High and Low Spin
Chemical Nomenclature
Electrochemical Conventions
Lewis Formulas (Structures)
Review of Oxidation and Reduction Reactions
Solving for the Cell Potential

Exam Reviews
Concepts for Midterm I
Concepts for Midterm II
Descriptive Chemistry Overview for Exam II
Extra Problems for Exam II
Solutions for Extra Problems for Exam II
Concepts for Final Exam
Extra Problems for Final

2 https://chem.libretexts.org/@go/page/182232
Licensing
A detailed breakdown of this resource's licensing can be found in Back Matter/Detailed Licensing.

1 https://chem.libretexts.org/@go/page/428097
CHAPTER OVERVIEW
Front Matter
TitlePage
InfoPage
Table of Contents
Licensing

1
UC Davis: Chem 2C General Chemistry iiI

LibreTexts
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 MindTouch® 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. Unless otherwise noted, LibreTexts content is licensed by CC BY-NC-SA 3.0.
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 03/12/2023


TABLE OF CONTENTS

AGENDA
Grading on a Curve

Text
Licensing
1: Chemistry Primer
1.1: Oxidation-Reduction Reactions
1.2: Balancing Redox Reactions
1.3: Free Energy and Equilibrium (quick)
1.4: Free Energy and Equilibrium (complete)
1.5: Electronic Con gurations - Spin Quantum Number
1.6: Electronic Con gurations - Hund's Rule
1.7: Electronic Con gurations - The Aufbau Process
1.8: Electronic Con gurations - Pauli Exclusion Principle
2: Electrochemistry
2.1: Galvanic Cells
2.2: Standard Electrode Potentials
2.3: Strengths of Oxidants and Reductants
2.4: E, ΔG, and K (The Big Triangle of Chemistry)
2.5: Cell Potential as a Function of Concentrations
2.6: Batteries- Producing Electricity Through Chemical Reactions
2.7: Corrosion- Unwanted Voltaic Cells
2.8: Electrolysis- Causing Nonspontaneous Reactions to Occur
3: Coordination Chemistry
3.1: Werner’s Theory of Coordination Compounds
3.2: Ligands
3.3: Nomenclature
3.4: Isomerism
3.5: Optical Isomerism in Metal Complexes
3.6: Stability Aspects of Complex-Ion Equilibria
3.7: Chelation
3.8: Bonding in Octahedral Complex Ions- Crystal Field Theory
3.9: Bonding in Non-Octahedral Complex Ions- Crystal Field Theory
3.10: Magnetic Behavior of Atoms, Molecules, and Materials
3.11: Magnetic Behavior of Complex Ions
3.12: Optical Properties of Coordination Compounds (Color)
4: Descriptive Chemistry
5: Chemical Kinetics
5.1: Factors that Affect Reaction Rates
5.2: Reaction Rates
5.3: Concentration and Rates (Rate Laws)
5.4: Determining Rate Laws from Initial Rates (Differential Rate Laws)
5.5: The Change of Concentration with Time (Integrated Rate Laws)
5.6: Using Graphs to Determine (Integrated) Rate Laws
5.7: Collision Theory

1 https://chem.libretexts.org/@go/page/182232
5.8: Temperature and Rate
5.9: Reaction Mechanisms
5.10: Catalysis
5.11: Exercises
5.12: Chemical Kinetics (Summary)
6: Nuclear Chemistry
6.1: Components of the Nucleus
6.2: Nuclear Reactions
6.3: Nuclear Radiation
6.4: Rates of Radioactive Decay
6.5: Stability of the Atomic Nucleus
6.6: The Origin of the Elements
6.7: Transmutation of the Elements
6.8: Nuclear Fission
6.9: Nuclear Fusion
6.10: Nuclear Chemistry (Exercises)

Worksheets
1: Balancing Redox Reactions (Worksheet)
2: Galvanic Cells (Worksheet)
3: Complex Ions and Nomenclature (Worksheet)
4: Complex Ion Structure and Geometry (Worksheet)
5: Crystal Field Theory (Worksheet)
6: Magnetism and Colors in Coordination Complexes (Worksheet)
7: Kinetics I (Worksheet)
8: Kinetics II (Worksheet)
9: Kinetics III (Worksheet)
10: Fundamentals of Nuclear Chemistry (WorkSheet)
11: Transmutation and Nuclear Kinetics (Worksheet)

Student Academic Success Center Workshops


Additional Resources: High and Low Spin
Chemical Nomenclature
Electrochemical Conventions
Lewis Formulas (Structures)
Review of Oxidation and Reduction Reactions
Solving for the Cell Potential

Exam Reviews
Concepts for Midterm I
Concepts for Midterm II
Descriptive Chemistry Overview for Exam II
Extra Problems for Exam II
Solutions for Extra Problems for Exam II
Concepts for Final Exam
Extra Problems for Final

2 https://chem.libretexts.org/@go/page/182232
Licensing
A detailed breakdown of this resource's licensing can be found in Back Matter/Detailed Licensing.

1 https://chem.libretexts.org/@go/page/428097
CHAPTER OVERVIEW
1: Chemistry Primer
1.1: Oxidation-Reduction Reactions
1.2: Balancing Redox Reactions
1.3: Free Energy and Equilibrium (quick)
1.4: Free Energy and Equilibrium (complete)
1.5: Electronic Configurations - Spin Quantum Number
1.6: Electronic Configurations - Hund's Rule
1.7: Electronic Configurations - The Aufbau Process
1.8: Electronic Configurations - Pauli Exclusion Principle

1: Chemistry Primer is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

1
1.1: Oxidation-Reduction Reactions
An oxidation-reduction (redox) reaction is a type of chemical reaction that involves a transfer of electrons between two species. An
oxidation-reduction reaction is any chemical reaction in which the oxidation number of a molecule, atom, or ion changes by
gaining or losing an electron. Redox reactions are common and vital to some of the basic functions of life, including
photosynthesis, respiration, combustion, and corrosion or rusting.

Rules for Assigning Oxidation States


The oxidation state (OS) of an element corresponds to the number of electrons, e-, that an atom loses, gains, or appears to use
when joining with other atoms in compounds. In determining the oxidation state of an atom, there are seven guidelines to
follow:
1. The oxidation state of an individual atom is 0.
2. The total oxidation state of all atoms in: a neutral species is 0 and in an ion is equal to the ion charge.
3. Group 1 metals have an oxidation state of +1 and Group 2 an oxidation state of +2
4. The oxidation state of fluorine is -1 in compounds
5. Hydrogen generally has an oxidation state of +1 in compounds
6. Oxygen generally has an oxidation state of -2 in compounds
7. In binary metal compounds, Group 17 elements have an oxidation state of -1, Group 16 elements of -2, and Group 15
elements of -3.

The sum of the oxidation states is equal to zero for neutral compounds and equal to the
charge for polyatomic ion species.
Example 1.1.1: Assigning Oxidation States

Determine the Oxidation States of each element in the following reactions:


a. Fe(s) + O 2
(g) ⟶ Fe O (g)
2 3

b. Fe (aq)
2 +

c. Ag(s) + H 2
S ⟶ Ag S(g) + H (g)
2 2

Solutions
A. Fe and O are free elements; therefore, they each have an oxidation state of 0 according to Rule #1. The product has a total
2

oxidation state equal to 0, and following Rule #6, O has an oxidation state of -2, which means Fe has an oxidation state of
+3.
B. The oxidation state of Fe ions just corresponds to its charge since it is a single element species; therefore, the oxidation
state is +2.
C. Ag has an oxidation state of 0, H has an oxidation state of +1 according to Rule #5, H has an oxidation state of 0, S has an
2

oxidation state of -2 according to Rule #7, and hence Ag in Ag S has an oxidation state of +1.
2

Example 1.1.2: Assigning Oxidation States

Determine the oxidation states of the phosphorus atom bold element in each of the following species:
a. Na PO3 3

b. H PO
2

Solutions
a. The oxidation numbers of Na and O are +1 and -2. Because sodium phosphite is neutral species, the sum of the oxidation
numbers must be zero. Letting x be the oxidation number of phosphorus, 0= 3(+1) + x + 3(-2). x=oxidation number of P=
+3.
b. Hydrogen and oxygen have oxidation numbers of +1 and -2. The ion has a charge of -1, so the sum of the oxidation
numbers must be -1. Letting y be the oxidation number of phosphorus, -1= y + 2(+1) +4(-2), y= oxidation number of P= +5.

1.1.1 https://chem.libretexts.org/@go/page/13350
Example 1.1.3: Identifying Reduced and Oxidized Elements

Determine which element is oxidized and which element is reduced in the following reactions (be sure to include the oxidation
state of each):
a. Zn + 2 H ⟶ Zn + H
+ 2 +
2

b. 2 Al + 3 Cu ⟶ 2 Al + 3 Cu
2 + 3 +

c. CO + 2 H ⟶ CO + H O
2 −

3
+

2 2

Solutions
a. Zn is oxidized (Oxidation number: 0 → +2); H+ is reduced (Oxidation number: +1 → 0)
b. Al is oxidized (Oxidation number: 0 → +3); Cu2+ is reduced (+2 → 0)
c. This is not a redox reaction because each element has the same oxidation number in both reactants and products: O= -2, H=
+1, C= +4.

An atom is oxidized if its oxidation number increases, the reducing agent, and an atom is reduced if its oxidation number decreases,
the oxidizing agent. The atom that is oxidized is the reducing agent, and the atom that is reduced is the oxidizing agent. (Note: the
oxidizing and reducing agents can be the same element or compound).

Oxidation-Reduction Reactions
Redox reactions are comprised of two parts, a reduced half and an oxidized half, that always occur together. The reduced half gains
electrons and the oxidation number decreases, while the oxidized half loses electrons and the oxidation number increases. Simple
ways to remember this include the mnemonic devices OIL RIG, meaning "oxidation is loss" and "reduction is gain." There is no
net change in the number of electrons in a redox reaction. Those given off in the oxidation half reaction are taken up by another
species in the reduction half reaction.
The two species that exchange electrons in a redox reaction are given special names:
1. The ion or molecule that accepts electrons is called the oxidizing agent - by accepting electrons it oxidizes other species.
2. The ion or molecule that donates electrons is called the reducing agent - by giving electrons it reduces the other species.
Hence, what is oxidized is the reducing agent and what is reduced is the oxidizing agent. (Note: the oxidizing and reducing agents
can be the same element or compound, as in disproportionation reactions discussed below).
A good example of a redox reaction is the thermite reaction, in which iron atoms in ferric oxide lose (or give up) O atoms to Al

atoms, producing Al O . 2 3

Fe O (s) + 2 Al(s) → Al O (s) + 2 Fe(l)


2 3 2 3

Example 1.1.4: Identifying Oxidizing and Reducing Agents

Determine what is the oxidizing and reducing agents in the following reaction.
+ 2 +
Zn + 2 H ⟶ Zn +H
2

Solution
The oxidation state of H changes from +1 to 0, and the oxidation state of Zn changes from 0 to +2. Hence, Zn is oxidized and
acts as the reducing agent. H ion is reduced and acts as the oxidizing agent.
+

Combination Reactions
Combination reactions are among the simplest redox reactions and, as the name suggests, involves "combining" elements to form
a chemical compound. As usual, oxidation and reduction occur together. The general equation for a combination reaction is given
below:

A + B ⟶ AB

1.1.2 https://chem.libretexts.org/@go/page/13350
Example 1.1.5: Combination Reaction

Consider the combination reaction of hydrogen and oxygen

H +O ⟶ H O
2 2 2

Solution
0 + 0 → (2)(+1) + (-2) = 0
In this reaction both H2 and O2 are free elements; following Rule #1, their oxidation states are 0. The product is H2O, which
has a total oxidation state of 0. According to Rule #6, the oxidation state of oxygen is usually -2. Therefore, the oxidation state
of H in H2O must be +1.

Decomposition Reactions
A decomposition reaction is the reverse of a combination reaction, the breakdown of a chemical compound into individual
elements:

AB ⟶ A + B

Example 1.1.6: Decomposition Reaction

Consider the following reaction:

H O⟶ H +O
2 2 2

This follows the definition of the decomposition reaction, where water is "decomposed" into hydrogen and oxygen.
(2)(+1) + (-2) = 0 → 0 + 0
As in the previous example the H O has a total oxidation state of 0; thus, according to Rule #6 the oxidation state of oxygen is
2

usually -2, so the oxidation state of hydrogen in H O must be +1.


2

Note that the autoionization reaction of water is not a redox nor decomposition reaction since the oxidation states do not
change for any element:
+ −
H O⟶ H + OH
2

Single Replacement Reactions


A single replacement reaction involves the "replacing" of an element in the reactants with another element in the products:

A + BC ⟶ AB + C

Example 1.1.7: Single Replacement Reaction

Equation:

Cl + Na Br → Na Cl + Br
2 ––– 2
––

Calculation:
(0) + ((+1) + (-1) = 0) -> ((+1) + (-1) = 0) + 0
In this equation, Br is replaced with Cl, and the Cl atoms in Cl are reduced, while the Br ion in NaBr is oxidized.
2

Double Replacement Reactions


A double replacement reaction is similar to a single replacement reaction, but involves "replacing" two elements in the reactants,
with two in the products:

AB + CD ⟶ AD + CB

An example of a double replacement reaction is the reaction of magnesium sulfate with sodium oxalate

1.1.3 https://chem.libretexts.org/@go/page/13350
MgSO (aq) + Na C O (aq) ⟶ MgC O (s) + Na SO (aq)
4 2 2 4 2 4 2 4

Combustion Reactions
Combustion is the formal terms for "burning" and typically involves a substance reacts with oxygen to transfer energy to the
surroundings as light and heat. Hence, combustion reactions are almost always exothermic. For example, internal combustion
engines rely on the combustion of organic hydrocarbons C H to generate CO and H O :
x y 2 2

Cx Hy + O ⟶ CO +H O
2 2 2

Although combustion reactions typically involve redox reactions with a chemical being oxidized by oxygen, many chemicals can
"burn" in other environments. For example, both titanium and magnesium metals can burn in nitrogen as well:

2 Ti(s) + N (g) ⟶ 2 TiN(s)


2

3 Mg(s) + N (g) ⟶ Mg N (s)


2 3 2

Moreover, chemicals can be oxidized by other chemicals than oxygen, such as Cl


2
or F
2
; these processes are also considered
combustion reactions.

Example 1.1.8: Identifying Combustion Reactions

Which of the following are combustion reactions?


a. 2 H O → 2 H + O
2 2 2

b. 4 Fe + 3 O → 2 Fe O
2 2 3

c. 2 AgNO + H S → Ag
3 2 2
S + 2 NHO
3

d. 2 Al + N → 2 AlN
2 4

Solution
Both reaction b and reaction d are combustion reactions, although with different oxidizing agents. Reaction b is the
conventional combustion reaction using O and reaction uses N instead.
2 2

Disproportionation Reactions
In disproportionation reactions, a single substance can be both oxidized and reduced. These are known as disproportionation
reactions, with the following general equation:
+n −n
2A ⟶ A +A

where n is the number of electrons transferred. Disproportionation reactions do not need begin with neutral molecules, and can
involve more than two species with differing oxidation states (but rarely).

Example 1.1.9: Disproportionation Reaction

Disproportionation reactions have some practical significance in everyday life, including the reaction of hydrogen peroxide,
H O
2 2
poured over a cut. This a decomposition reaction of hydrogen peroxide, which produces oxygen and water. Oxygen is
present in all parts of the chemical equation and as a result it is both oxidized and reduced. The reaction is as follows:

2 H O (aq) ⟶ 2 H O(l) + O (g)


2 2 2 2

Dicussion
On the reactant side, H has an oxidation state of +1 and O has an oxidation state of -1, which changes to -2 for the product
H O (oxygen is reduced), and 0 in the product O (oxygen is oxidized).
2 2

Exercise 1.1.9

Which element undergoes a bifurcation of oxidation states in this disproportionation reaction:

HNO ⟶ HNO + NO + H O
2 3 2

1.1.4 https://chem.libretexts.org/@go/page/13350
Answer
The N atom undergoes disproportionation. You can confirm that by identifying the oxidation states of each atom in each
species.

References
1. Petrucci, et al. General Chemistry: Principles & Modern Applications. 9th ed. Upper Saddle River, New Jersey:
Pearson/Prentice Hall, 2007.
2. Sadava, et al. Life: The Science of Biology. 8th ed. New York, NY. W.H. Freeman and Company, 2007

Contributors and Attributions


Christopher Spohrer (UCD), Christina Breitenbuecher (UCD), Luvleen Brar (UCD)

1.1: Oxidation-Reduction Reactions is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

1.1.5 https://chem.libretexts.org/@go/page/13350
1.2: Balancing Redox Reactions
Oxidation-Reduction Reactions, or redox reactions, are reactions in which one reactant is oxidized and one reactant is reduced
simultaneously. This module demonstrates how to balance various redox equations.

Identifying Redox Reactions


The first step in balancing any redox reaction is determining whether or not it is even an oxidation-reduction reaction. This requires that
one and typically more species changing oxidation states during the reaction. To maintain charge neutrality in the sample, the redox
reaction will entail both a reduction component and an oxidation components. These are often separated into independent two hypothetical
half-reactions to aid in understanding the reaction. This requires identifying which element is oxidized and which element is reduced. For
example, consider this reaction:
+ 2 +
Cu(s) + 2 Ag (aq) → Cu (aq) + 2 Ag(s)

The first step in determining whether the reaction is a redox reaction is to split the equation into two hypothetical half-reactions. Let's start
with the half-reaction involving the copper atoms:
2 +
Cu(s) → Cu (aq)

The oxidation state of copper on the left side is 0 because it is an element on its own. The oxidation state of copper on the right hand side
of the equation is +2. The copper in this half-reaction is oxidized as the oxidation states increases from 0 in Cu to +2 in Cu . Now 2 +

consider the silver atoms


+
2 Ag (aq) → 2 Ag(s)

In this half-reaction, the oxidation state of silver on the left side is a +1. The oxidation state of silver on the right is 0 because it is an pure
element. Because the oxidation state of silver decreases from +1 to 0, this is the reduction half-reaction.
Consequently, this reaction is a redox reaction as both reduction and oxidation half-reactions occur (via the transfer of electrons, that are
not explicitly shown in equations 2). Once confirmed, it often necessary to balance the reaction (the reaction in equation 1 is balanced
already though), which can be accomplished in two ways because the reaction could take place in neutral, acidic or basic conditions.

Balancing Redox Reactions


Balancing redox reactions is slightly more complex than balancing standard reactions, but still follows a relatively simple set of rules. One
major difference is the necessity to know the half-reactions of the involved reactants; a half-reaction table is very useful for this. Half-
reactions are often useful in that two half reactions can be added to get a total net equation. Although the half-reactions must be known to
complete a redox reaction, it is often possible to figure them out without having to use a half-reaction table. This is demonstrated in the
acidic and basic solution examples. Besides the general rules for neutral conditions, additional rules must be applied for aqueous reactions
in acidic or basic conditions.
One method used to balance redox reactions is called the Half-Equation Method. In this method, the equation is separated into two half-
equations; one for oxidation and one for reduction.

Half-Equation Method to Balance redox Reactions in Acidic Aqueous Solutions

Each reaction is balanced by adjusting coefficients and adding H 2


O , H , and e in this order:
+ −

1. Balance elements in the equation other than O and H .


2. Balance the oxygen atoms by adding the appropriate number of water (H O ) molecules to the opposite side of the equation.
2

3. Balance the hydrogen atoms (including those added in step 2 to balance the oxygen atom) by adding H ions to the opposite side
+

of the equation.
4. Add up the charges on each side. Make them equal by adding enough electrons (e ) to the more positive side. (Rule of thumb:

e

and H are almost always on the same side.)
+

5. The e on each side must be made equal; if they are not equal, they must be multiplied by appropriate integers (the lowest

common multiple) to be made the same.


6. The half-equations are added together, canceling out the electrons to form one balanced equation. Common terms should also be
canceled out.
The equation can now be checked to make sure that it is balanced.

1.2.1 https://chem.libretexts.org/@go/page/13347
Half-Equation Method to Balance redox Reactions in Basic Aqueous Solutions
If the reaction is being balanced in a basic solution, the above steps are modified with the the addition of one step between #3 and #4:
3b Add the appropriate number of OH to neutralize all H and to convert into water molecules.
− +

The equation can now be checked to make sure that it is balanced.

Neutral Conditions
The first step to balance any redox reaction is to separate the reaction into half-reactions. The substance being reduced will have electrons
as reactants, and the oxidized substance will have electrons as products. (Usually all reactions are written as reduction reactions in half-
reaction tables. To switch to oxidation, the whole equation is reversed and the voltage is multiplied by -1.) Sometimes it is necessary to
determine which half-reaction will be oxidized and which will be reduced. In this case, whichever half-reaction has a higher reduction
potential will by reduced and the other oxidized.

Example 1.2.1: Balancing in a Neutral Solution

Balance the following reaction


+ 3 +
Cu (aq) + Fe(s) → Fe (aq) + Cu(s)

Solution
Step 1: Separate the half-reactions. By searching for the reduction potential, one can find two separate reactions:
+ −
Cu (aq) + e → Cu(s)

and
3 + −
Fe (aq) + 3 e → Fe(s)

The copper reaction has a higher potential and thus is being reduced. Iron is being oxidized so the half-reaction should be flipped.
This yields:
+ −
Cu (aq) + e → Cu(s)

and
3 + −
Fe(s) → Fe (aq) + 3 e

Step 2: Balance the electrons in the equations. In this case, the electrons are simply balanced by multiplying the entire
→ C u(s) half-reaction by 3 and leaving the other half reaction as it is. This gives:
+ −
C u (aq) + e

+ −
3 Cu (aq) + 3 e → 3 Cu(s)

and
3 + −
Fe(s) → Fe (aq) + 3 e

Step 3: Adding the equations give:


+ − 3 + −
3 Cu (aq) + 3 e + Fe(s) → 3 Cu(s) + Fe (aq) + 3 e

The electrons cancel out and the balanced equation is left.


+ 3 +
3 Cu (aq) + Fe(s) → 3 Cu(s) + Fe (aq)

Acidic Conditions
Acidic conditions usually implies a solution with an excess of H concentration, hence making the solution acidic. The balancing starts by
+

separating the reaction into half-reactions. However, instead of immediately balancing the electrons, balance all the elements in the half-
reactions that are not hydrogen and oxygen. Then, add H O molecules to balance any oxygen atoms. Next, balance the hydrogen atoms
2

by adding protons (H ). Now, balance the charge by adding electrons and scale the electrons (multiply by the lowest common multiple)
+

so that they will cancel out when added together. Finally, add the two half-reactions and cancel out common terms.

1.2.2 https://chem.libretexts.org/@go/page/13347
Example 1.2.2: Balancing in a Acid Solution

Balance the following redox reaction in acidic conditions.


2 − 3 + −
Cr O (aq) + HNO (aq) → Cr (aq) + NO (aq)
2 7 2 3

Solution
Step 1: Separate the half-reactions. The table provided does not have acidic or basic half-reactions, so just write out what is known.
2 − 3 +
Cr O (aq) → Cr (aq)
2 7


HNO (aq) → NO (aq)
2 3

Step 2: Balance elements other than O and H. In this example, only chromium needs to be balanced. This gives:
2 − 3 +
Cr O7 (aq) → 2 Cr (aq)
2

and

HNO (aq) → NO3 (aq)
2

Step 3: Add H2O to balance oxygen. The chromium reaction needs to be balanced by adding 7 H O
2
molecules. The other reaction
also needs to be balanced by adding one water molecule. This yields:
2 − 3 +
Cr O (aq) → 2 Cr (aq) + 7 H O(l)
2 7 2

and

HNO (aq) + H O(l) → NO (aq)
2 2 3

Step 4: Balance hydrogen by adding protons (H+). 14 protons need to be added to the left side of the chromium reaction to balance
the 14 (2 per water molecule * 7 water molecules) hydrogens. 3 protons need to be added to the right side of the other reaction.
+ 2 − 3 +
14 H (aq) + Cr O (aq) → 2 Cr (aq) + 7 H O(l)
2 7 2

and
+ −
HNO (aq) + H O(l) → 3 H (aq) + NO3 (aq)
2 2

Step 5: Balance the charge of each equation with electrons. The chromium reaction has (14+) + (2-) = 12+ on the left side and (2 *
3+) = 6+ on the right side. To balance, add 6 electrons (each with a charge of -1) to the left side:
− + 2 − 3 +
6e + 14 H (aq) + Cr O (aq) → 2 Cr (aq) + 7 H O(l)
2 7 2

For the other reaction, there is no charge on the left and a (3+) + (-1) = 2+ charge on the right. So add 2 electrons to the right side:
+ − −
HNO (aq) + H O(l) → 3 H (aq) + NO (aq) + 2 e
2 2 3

Step 6: Scale the reactions so that the electrons are equal. The chromium reaction has 6e- and the other reaction has 2e-, so it should
be multiplied by 3. This gives:
− + 2 − 3 +
6e + 14 H (aq) + Cr O (aq) → 2 Cr (aq) + 7 H O(l)⋅
2 7 2

and
+ − −
3 × [ HNO (aq) + H O(l) → 3H (aq) + NO (aq) + 2 e ]
2 2 3

+ − −
3 HNO (aq) + 3 H O(l) → 9 H (aq) + 3 NO (aq) + 6 e
2 2 3

Step 7: Add the reactions and cancel out common terms.


+ − −
3 HNO (aq) + 3 H O(l) → 9 H (aq) + 3 NO3 (aq) + 6 e
2 2

− + 2 − 3 +
6e + 14 H (aq) + Cr O (aq) → 2 Cr (aq) + 7 H O(l)
2 7 2

4
− + 2 − + − − 3 +
3 HNO (aq) + 3 H O(l) + 6e + 14 H (aq) + Cr O (aq) → 9H (aq) + 3 NO (aq) + 6e + 2 Cr (aq) + 7 H O(l)
2 2 2 7 3 2

The electrons cancel out as well as 3 water molecules and 9 protons. This leaves the balanced net reaction of:

1.2.3 https://chem.libretexts.org/@go/page/13347
+ 2 − − 3 +
3 HNO (aq) + 5 H (aq) + Cr O (aq) → 3 NO3 (aq) + 2 Cr (aq) + 4 H O(l)
2 2 7 2

Basic Conditions
Bases dissolve into OH ions in solution; hence, balancing redox reactions in basic conditions requires OH . Follow the same steps as
− −

for acidic conditions. The only difference is adding hydroxide ions to each side of the net reaction to balance any H . OH and H ions + − +

on the same side of a reaction should be added together to form water. Again, any common terms can be canceled out.

Example 1.2.1: Balancing in Basic Solution

Balance the following redox reaction in basic conditions.


2 +
Ag(s) + Zn (aq) → Ag O(aq) + Zn(s)
2

Solution
Go through all the same steps as if it was in acidic conditions.
Step 1: Separate the half-reactions.

Ag(s) → Ag O(aq)
2

2 +
Zn (aq) → Zn(s)

Step 2: Balance elements other than O and H.

2 Ag(s) → Ag O(aq)
2

2 +
Zn (aq) → Zn(s)

Step 3: Add H2O to balance oxygen.

H O(l) + 2 Ag(s) → Ag O(aq)


2 2

2 +
Zn (aq) → Zn(s)

Step 4: Balance hydrogen with protons.


+
H O(l) + 2 Ag(s) → Ag O(aq) + 2 H (aq)
2 2

2 +
Zn (aq) → Zn(s)

Step 5: Balance the charge with e-.


+ −
H O(l) + 2 Ag(s) → Ag O(aq) + 2 H (aq) + 2 e
2 2

2 + −
Zn (aq) + 2 e → Zn(s)

Step 6: Scale the reactions so that they have an equal amount of electrons. In this case, it is already done.
Step 7: Add the reactions and cancel the electrons.
2 + +
H O(l) + 2 Ag(s) + Zn (aq) → Zn(s) + Ag O(aq) + 2 H (aq)⋅
2 2

- + -
Step 8: Add OH to balance H . There are 2 net protons in this equation, so add 2 OH ions to each side.
2 + − + −
H O(l) + 2 Ag(s) + Zn (aq) + 2 OH (aq) → Zn(s) + Ag O(aq) + 2 H (aq) + 2 OH (aq)⋅
2 2

Step 9: Combine OH- ions and H+ ions that are present on the same side to form water.
2 + −
H O(l) + 2 Ag(s) + Zn (aq) + 2 OH (aq) ⟶ Zn(s) + Ag O(aq) + 2 H O(l)
2 2 2

Step 10: Cancel common terms.


2 + −
2 Ag(s) + Zn (aq) + 2 OH (aq) → Zn(s) + Ag O(aq) + H O(l)
2 2

1.2.4 https://chem.libretexts.org/@go/page/13347
References
1. Petrucci, Ralph, William Harwood, Geoffrey Herring, and Jeffry Madura. General Chemistry: Principles & Modern Applications. 9th
edition. Upper Saddle River, New Jersey: Pearson Prentince Hall, 2007.
2. Helmenstine, Anne Marie. "How to Balance Redox Reactions - Balancing Redox Reactions." Balancing Redox Reactions - Half-
Reaction Method (2009): n. pag. Web. 1 Dec 2009. http://chemistry.about.com/od/genera...s/redoxbal.htm
3. Stanitski, Conrad L. "Chemical Equations." Chemistry Explained Foundations and Applications. 1st. Chemistry Encyclopedia, 2009.
Print.
4. "How to Balance Redox Equations." Youtube. Web. 1 Dec 2009.

1.2: Balancing Redox Reactions is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

1.2.5 https://chem.libretexts.org/@go/page/13347
1.3: Free Energy and Equilibrium (quick)
The balance between reactants and products in a reaction will be determined by the free energy difference between the two sides of
the reaction. The greater the free energy difference, the more the reaction will favor one side or the other. The smaller the free
energy difference, the closer the mixture will get to equal parts reactants and products (loosely speaking).
Exactly where the balance lies in an equilibrium reaction is described by the equilibrium constant. The equilibrium constant is just
the ratio of products to reactants, once the reaction has settled to equilibrium. That's the point at which the forward and reverse
reactions are balanced, so that the ratio of products to reactants is unchanging.
A reaction has reached equilibrium when the reaction has stopped progressing (i.e., no change in concentrations although at a
microscopic level both forward and reverse reactions occur), so that the amount of reactants that have turned into products
remains constant, and the amount of reactants left over stays constant.
The equilibrium constant is the ratio of products to reactants when the reaction has reached equilibrium.
The equilibrium constant could be a large number (like a thousand). That means that there are much more products than reactants at
equilibrium. It could also be a very small fraction (like one millionth). That would indicate that the reaction does not proceed very
far, producing only a tiny amount of products at equilibrium.
Every reaction has an equilibrium constant
A very large equilibrium constant (in the millions or billions) means the reaction goes "to completion", with all reactants
essentially converted into products
A tiny equilibrium (very close to zero) constant means the reaction hardly moves forward at all.
A modest equilibrium constant (close to one, or as close to one as numbers like 0.01 or 100) is considered to be a true
equilibrium reaction, in which there is a significant amount of both products and reactants.
The equilibrium constant is related to the free energy change of the reaction by the expression:
−ΔG/RT
K =e (1.3.1)

or
ΔG
ln K = − (1.3.2)
RT

in which T is the temperature in Kelvin and R is the "gas constant" (1.986 cal/K mol). Remember, e is just a number that occurs
frequently in mathematical relationships in nature (sort of like π); it has a value of about 2.718. This expression for K does make
some assumptions about the conditions that we won't worry about; we are using a slightly simplified model.

Relating Gibbs energy and the equilibrium


Let's look at the form of this relationship between free energy and the equilibrium constant. First, we will see how we deal with
endergonic versus exergonic reactions. The free energy changes in opposite directions in these two cases, and we usually deal with
opposites by giving one quantity a positive sign and one quantity a negative sign. A reaction in which the free energy increases is
given a positive value for its free energy. On the other hand, if free energy decreases over the course of the reaction, we show that
by using a negative number for the value of the free energy.
If ΔG is negative, the exponent in the relationship becomes positive (because it is multiplied by -1 in the expression). Since e to a
positive power will usually be a number greater than one, the relationship suggests there are more products than reactants. That's
good, because the reaction is exergonic, and we expect the reaction to go forward. What's more, the larger the value of ΔG, the
more product-favored the reaction will be.
10large number is a large number.
10small number is a smaller number.
However, if ΔG is a positive number, then the exponent in the relationship becomes negative. An number with a negative exponent,
by the rules of exponents, is the same as the inverse of the number with a positive exponent of the same size.
In other words, 10-2 = 1 / 102.
10negative number is a fraction.

1.3.1 https://chem.libretexts.org/@go/page/13349
That means if ΔG is positive, the equilibrium constant becomes a fraction. That's because that positive value of ΔG is multiplied by
-1 in the expression, becoming negative, and then it's placed in the exponent. That's good, because a positive value of ΔG
corresponds to an endergonic reaction, which does not favor product formation.

Other factors
There are other factors in the expression relating ΔG to the equilibrium constant. One of them, R, is just a "fudge factor"; it's the
number that, when placed in the expression, makes the relationship agree with reality.
Moreover, it is a constant, so it does not change. However, the other factor is temperature, which does change. That means that the
equilibrium constant may change with different temperatures. Overall, the effect of temperature is to make the exponent in the
expression a smaller number. That's because the free energy is divided by the temperature and the gas constant; the resulting
number becomes the exponent in the relationship. At the extreme, a high temperature could make the exponent into a very, very
small number, something close to zero. What happens then?
100 = 1
e0 = 1
As the exponent gets smaller and smaller, the equilibrium constant could approach 1. That means there would be more or less equal
amounts of products and reactants in our simplified approach.
However, the fact that there is a temperature factor in the expression for ΔG itself means that there is a limit to how small K will
get as the temperature increases. At some point, the two values for temperature cancel out altogether and the expression becomes K
= e (ΔS/R). At that point, the equilibrium constant is independent of temperature and is based only on internal entropy differences
between the two sides of the reaction.
This relationship is useful because of its predictive value. Qualtitatively, it confirms ideas we had already developed about
thermodynamics.
Highly exergonic reactions (large, negative/decreasing ΔG) favor products.
Highly endergonic reactions (large, positive/increasing ΔG) favor reactants.
Reactions with small free energy changes lead to equilibrium mixtures of both products and reactants.

Problems
TD6.1. Arrange the following series of numbers from the largest quantity to the smallest, from left to right.
A. 105 104 106
B. 23 26 22
C. 33 30 32
D. e2 e1 e4
E. 10-1 10-5 10-3
F. 1 / 10 1 / 25 1 / 50
G. 20.5 20.1 20.9
Problem TD6.2. Given the following free energy differences, arrange the corresponding equilibrium constants from largest
to smallest.
A. 25 kcal/mol 17 kcal/mol 9 kcal/mol
B. 16 kcal/mol 19 kcal/mol 21 kcal/mol
C. 7 kcal/mol 22 kcal/mol 13 kcal/mol
D. -17 kcal/mol -3 kcal/mol -8 kcal/mol
E. -17 kcal/mol 3 kcal/mol -8 kcal/mol
Problem TD6.3: What is the value of the equilibrium constant at 300K in the following cases? (1 kcal = 1000 cal)
A. ΔG = 3 kcal /mol
B. ΔG = -2 kcal/mol
C. ΔG = -5 kcal /mol
D. ΔG = 15 kcal/mol
E. ΔG = -10 kcal/mol

1.3.2 https://chem.libretexts.org/@go/page/13349
F. The free energy increases by 8 kcal/mol over the reaction.
G. The free energy decreases by 1 kcal/mol over the reaction.
Problem TD6.4. In which of the cases in TD6.3. do you think there would be significant amounts of both products and
reactants at equilibrium?
Problem TD6.5. The mathematical expression for the equilibrium constant says that K will get smaller at higher
temperatures. Explain this phenomenon without the mathematical expression in terms of what you know about
temperature and energy.

Contributors and Attributions


Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

1.3: Free Energy and Equilibrium (quick) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
Free Energy and Equilibrium by Chris Schaller is licensed CC BY-NC 3.0.

1.3.3 https://chem.libretexts.org/@go/page/13349
1.4: Free Energy and Equilibrium (complete)
 Learning Objectives

You are expected to be able to define and explain the significance of terms identified in italic type.
1. As a homogeneous chemical reaction proceeds, the Gibbs energies of the reactants become more negative and those of the
products more positive as the composition of the system changes.
2. The total Gibbs energy of the system (reactants + products) always becomes more negative as the reaction proceeds.
Eventually it reaches a minimum value at a system composition that defines the equilibrium composition of the system,
after which time no further net change will occur.
3. The equilibrium constant for the reaction is determined the standard Gibbs energy change:
ΔG° = -RT ln Kp
4. The sign of the temperature dependence of the equilibrium constant is governed by the sign of ΔH°. This is the basis of the
Le Chatelier Principle.
5. The Gibbs energies of solid and liquid components are constants that do not change with composition. Thus in
heterogeneous reactions such as phase changes, the total Gibbs energy does not pass through a minimum and when the
system is not at equilibrium only all-products or all-reactants will be stable.
6. Two reactions are coupled when the product of one reaction is consumed in the other. If ΔG° for the first reaction is
positive, the overall process can still be spontaneous if ΔG° for the second reaction is sufficiently negative— in which case
the second reaction is said to "drive" the first reaction.

Under conditions of constant temperature and pressure, chemical change will tend to occur in whatever direction leads to a
decrease in the value of the Gibbs energy. In this lesson we will see how G varies with the composition of the system as reactants
change into products. When G falls as far as it can, all net change comes to a stop. The equilibrium composition of the mixture is
determined by ΔG° which also defines the equilibrium constant K.

The Road to Equilibrium is Down the Gibbs Energy Hill


This means, of course, that if the total Gibbs energy G of a mixture of reactants and products goes through a minimum value as the
composition changes, then all net change will cease— the reaction system will be in a state of chemical equilibrium. You will recall
that the relative concentrations of reactants and products in the equilibrium state is expressed by the equilibrium constant. In this
lesson we will examine the relation between the Gibbs energy change for a reaction and the equilibrium constant.
To keep things as simple as possible, we will consider a homogeneous chemical reaction of the form
A+B ⇌ C +D (1.4.1)

in which all components are gases at the temperature of interest. If the sum of the standard Gibbs energies of the products is less
than that of the reactants, ΔG° for the reaction will be negative and the reaction will proceed to the right. But how far? If the
reactants are completely transformed into products, the equilibrium constant would be infinity. The equilibrium constants we
actually observe all have finite values, implying that even if the products have a lower Gibbs energy than the reactants, some of the
latter will always remain when the process comes to equilibrium.

A homogeneous reaction is one in which everything takes place in a single gas or liquid
phase.
To understand how equilibrium constants relate to ΔG° values, assume that all of the reactants are gases, so that the Gibbs energy
of gas A, for example, is given at all times by

GA = G + RT ln PA (1.4.2)
A

The Gibbs energy change for the reaction is sum of the Gibbs energies of the products, minus the sum of Gibbs energies of the
reactants:

ΔG = GC + GD – GA – GB (1.4.3)
 
products reactants

1.4.1 https://chem.libretexts.org/@go/page/13348
Using Equation 1.4.2 to expand each term on the right of Equation 1.4.3, we have
∘ ∘ ∘ ∘
ΔG = (G + RT ln PC ) + (G + RT ln PD )– (G + RT ln PB )– (G + RT ln P + A) (1.4.4)
C D B A

We can now express the G terms collectively as ΔG , and combine the logarithmic pressure terms into a single fraction
∘ ∘

PC PD
ΔG = ΔG° + RT ln( ) (1.4.5)
PA PB

which is more conveniently expressed in terms of the reaction quotient Q.



ΔG = ΔG + RT ln Q (1.4.6)

The Gibbs energy G is a quantity that becomes more negative during the course of any natural process. Thus as a chemical reaction
takes place, G only falls and will never become more positive. Eventually a point is reached where any further transformation of
reactants into products would cause G to increase. At this point G is at a minimum (see below), and no further net change can take
place; the reaction is then at equilibrium.

Although Equations 1.4.2 -1.4.6 are strictly correct only for perfect gases, we will see later
that equations of similar form can be applied to many liquid solutions by substituting
concentrations for pressures.

 Example 1.4.1: Dissociation of Dinitrogen Tetroxide

Consider the gas-phase dissociation reaction

N O → 2 NO
2 4 2

which is a simple example of the Gibbs energy relationships in a homogeneous reaction.

Figure 1.4.1 : Gibbs energy for the thermal dissociation reaction of dinitrogen tetroxide
The Gibbs energy of 1 mole of N2O4 (1) is smaller than that of 2 moles of NO2 (2) by 5.3 kJ; thus ΔG = +5.3 kJ for the
o

complete transformation of reactants into products. The straight diagonal line shows the Gibbs energy of all possible
compositions if the two gases were prevented from mixing. The red curved line show the Gibbs energy of the actual reaction
mixture. This passes through a minimum at (3) where 0.814 mol of N O are in equilibrium with 0.372 mol of N O . The
2 4 2

difference (4) corresponds to the Gibbs energy of mixing of reactants and products which always results in an equilibrium
mixture whose Gibbs energy is lower than that of either pure reactants or pure products. Thus some amount of reaction will
occur even if ΔG° for the process is positive.
What’s the difference between ΔG and ΔG°?
It’s very important to be aware of this distinction; that little ° symbol makes a world of difference! First, the standard Gibbs
energy change ΔG° has a single value for a particular reaction at a given temperature and pressure; this is the difference
∘ ∘
∑ G (products)– ∑ G (reactants) (1.4.7)
f f

1.4.2 https://chem.libretexts.org/@go/page/13348
that are tabulated in thermodynamic tables. It corresponds to the Gibbs energy change for a process that never really happens:
the complete transformation of pure N2O4 into pure NO2 at a constant pressure of 1 atm.

Figure 1.4.2 : Gibbs energy change for the thermal dissociation reaction of dinitrogen tetroxide
The other quantity ΔG, defined by Equation 1.4.6, represents the total Gibbs energies of all substances in the reaction mixture
at any particular system composition. In contrast to ΔG which is a constant for a given reaction, ΔG varies continuously as

the composition changes, finally reaching zero at equilibrium. ΔG is the “distance” (in Gibbs energy) from the equilibrium
state of a given reaction. Thus for the limiting cases of pure N O or NO (as far from the equilibrium state as the system can
2 4 2

be!),
2
[N O2 ]
Q = = ±∞ (1.4.8)
[ N2 O4 ]

which makes the logarithm in Equation 1.4.6, and thus the value of ΔG, approach the same asymptotic limits (1) or (2). As the
reaction proceeds in the appropriate direction ΔG approaches zero; once there (3), the system is at its equilibrium composition
and no further net change will occur.

 Example 1.4.2: Isomerization of Butane

The standard molar Gibbs energy change for this very simple reaction is –2.26 kJ, but mixing of the unreacted butane with the
product brings the Gibbs energy of the equilibrium mixture down to about –3.1 kJ mol–1 at the equilibrium composition
corresponding to 77 percent conversion.

Notice particularly that


The sum of the Gibbs energies of the two gases (n-butane and iso-butane) separately varies linearly with the composition of
the mixture (red line ).
The green curve adds the Gibbs energy of mixing to the above sum; its minimum defines the equilibrium composition.
As the composition approaches the equilibrium value , ΔG (which denotes how much farther the Gibbs energy of the
system can fall) approaches zero.
The detailed calculations that lead to the values shown above can be found here.

1.4.3 https://chem.libretexts.org/@go/page/13348
Why reactions lead to mixtures of reactants and products
We are now in a position to answer the question posed earlier: if ΔG° for a reaction is negative, meaning that the Gibbs energies of
the products are more negative than those of the reactants, why will some of the latter remain after equilibrium is reached? The
answer is that no matter how low the Gibbs energy of the products, the Gibbs energy of the system can be reduced even more by
allowing some of the products to be "contaminated" (i.e., diluted) by some reactants. Owing to the entropy associated with mixing
of reactants and products, no homogeneous reaction will be 100% complete. An interesting corollary of this is that any reaction for
which a balanced chemical equation can be written can in principle take place to some extent, however minute that might be.
Gibbs energies of mixing of products with reactants tend to be rather small, so for reactions having ΔG° values that are highly
negative or positive (±20 kJ mol–1, say), the equilibrium mixture will, for all practical purposes, be either [almost] "pure" reactants
or products.

The Equilibrium Constant


Now let us return to Equation 1.4.6 which we reproduce here:

ΔG = ΔG + RT ln Q (1.4.9)

As the reaction approaches equilibrium, ΔG becomes less negative and finally reaches zero. At equilibrium ΔG = 0 and Q = K ,
so we can write (must know this!)

ΔG =– RT ln Kp (1.4.10)

in which K , the equilibrium constant expressed in pressure units, is the special value of
p Q that corresponds to the equilibrium
composition.
This equation is one of the most important in chemistry because it relates the equilibrium composition of a chemical reaction
system to measurable physical properties of the reactants and products. If you know the entropies and the enthalpies of formation
of a set of substances, you can predict the equilibrium constant of any reaction involving these substances without the need to know
anything about the mechanism of the reaction.

Instead of writing Equation 1.4.10 in terms of Kp, we can use any of the other forms of
the equilibrium constant such as Kc (concentrations), Kx (mole fractions), Kn(numbers of
moles), etc. Remember, however, that for ionic solutions especially, only the Ka, in which
activities are used, will be strictly valid.
It is often useful to solve Equation 1.4.10 for the equilibrium constant, yielding
−ΔG
K = exp ( ) (1.4.11)
RT

Figure 1.4.3
This relation is most conveniently plotted against the logarithm of K as shown in Figure 1.4.3, where it can be represented as a
straight line that passes through the point (0,0).

1.4.4 https://chem.libretexts.org/@go/page/13348
 Example 1.4.3

Calculate the equilibrium constant for the reaction from the following thermodynamic data:
+ –
H −
(aq) + OH (aq) ↽⇀
− H O(l)
2

+
H (aq)


OH (aq)

H2 O(l)

ΔHf°, kJ mol–1
0
–230.0
–285.8
S°, J K–1 mol–1
0*
–10.9
70.0
* Note that the standard entropy of the hydrogen ion is zero by definition. This reflects the fact that it is impossible to carry out
thermodynamic studies on a single charged species. All ionic entropies are relative to that of H (aq), which explains why some +

values (as for aqueous hydroxide ion) are negative.

Solution
From the above data, we can evaluate the following quantities:
o o o
ΔH = ∑ ΔH (products) − ∑ ΔH (reactants)
f f

= (– 285.8) − (−230)

−1
=– 55.8 kJ mol

o o
ΔS = ∑ ΔS (products) − ∑ ΔS°(reactants)

= (70.0)– (– 10.9)

−1 −1
= +80.8 J K mol

The value of ΔG° at 298 K is


o o
ΔH – T ΔS = (– 55800)– (298)(80.8)

–1
=– 79900 J mol

From Equation 1.4.11 we have


– 79900
K = exp( )
8.314 × 298

32.2 –14
=e = 1.01 × 10

Equilibrium and Temperature


We have already discussed how changing the temperature will increase or decrease the tendency for a process to take place,
depending on the sign of ΔS°. This relation can be developed formally by differentiating the relation
∘ ∘ ∘
ΔG = ΔH – T ΔS (1.4.12)

with respect to the temperature:



d(−ΔG )

= −ΔS (1.4.13)
dT

1.4.5 https://chem.libretexts.org/@go/page/13348
Hence, the sign of the entropy change determines whether the reaction becomes more or less allowed as the temperature increases.
We often want to know how a change in the temperature will affect the value of an equilibrium constant whose value is known at
some fixed temperature. Suppose that the equilibrium constant has the value K at temperature T and we wish to estimate K at
1 1 2

temperature T . Expanding Equation 1.4.11 in terms of ΔH and ΔS , we obtain


2
∘ ∘

∘ ∘
– RT1 ln K1 = ΔH – T1 ΔS (1.4.14)

and
∘ ∘
– RT2 ln K2 = ΔH – T2 ΔS (1.4.15)

Dividing both sides by RT and subtracting, we obtain


∘ ∘
ΔH ΔH
ln K1 − ln K2 = − ( − ) (1.4.16)
RT1 RT2

Which is most conveniently expressed as the ratio



K1 ΔH 1 1
ln =− ( − ) (1.4.17)
K2 R T1 T2

This is its theoretical foundation of Le Chatelier's Principle with respect to the effect of the temperature on equilibrium:
if the reaction is exothermic ΔH < 0 , then increasing temperature will make the second exponential term smaller and K will

decrease. The equilibrium will then “shift to the left”.


If ΔH > 0 , then increasing T will make the exponent less negative and K will increase and the equilibrium will “shift to the

right”.
This is an extremely important relationship, but not just because of its use in calculating the temperature dependence of an
equilibrium constant. Even more important is its application in the “reverse” direction to experimentally determine ΔH° from two
values of the equilibrium constant measured at different temperatures. Direct calorimetric determinations of heats of reaction are
not easy to make; relatively few chemists have the equipment and experience required for this rather exacting task. Measurement of
an equilibrium constant is generally much easier, and often well within the capabilities of anyone who has had an introductory
Chemistry course. Once the value of ΔH° is determined it can be combined with the Gibbs energy change (from a single
observation of K, through Equation 1.4.11) to allow ΔS° to be calculated through Equation 1.4.13.

Equilibrium Without Mixing: it's all or nothing


You should now understand that for homogeneous reactions (those that take place entirely in the gas phase or in solution) the
equilibrium composition will never be 100% products, no matter how much lower their Gibbs energy relative to the reactants. As
was summarized in the N2O4-dissociation example discussed previously. This is due to "dilution" of the products by the reactants.
In heterogeneous reactions (those which involve more than one phase) this dilution, and the effects that flow from it, may not be
possible.
A particularly simple but important type of a heterogeneous process is phase change. Consider, for example, an equilibrium
mixture of ice and liquid water. The concentration of H2O in each phase is dependent only on the density of the phase; there is no
way that ice can be “diluted” with water, or vice versa. This means that at all temperatures other than the freezing point, the lowest
Gibbs energy state will be that corresponding to pure ice or pure liquid. Only at the freezing point, where the Gibbs energies of
water and ice are identical, can both phases coexist, and they may do so in any proportion.

Gibbs energy of the ice-water system


Only at 0°C can ice and liquid water coexist in any proportion. Note that in contrast to the homogeneous N2O4 example, there is no
Gibbs energy minimum at intermediate compositions.

1.4.6 https://chem.libretexts.org/@go/page/13348
Coupled Reactions
Two reactions are said to be coupled when the product of one of them is the reactant in the other:

A → B

and

B → C

If the standard Gibbs energy of the first reaction is positive but that of the second reaction is sufficiently negative, then for the
overall process will be negative and we say that the first reaction is “driven” by the second one. This, of course, is just another way
of describing an effect that you already know as the Le Chatelier principle: the removal of substance B by the second reaction
causes the equilibrium of the first to “shift to the right”. Similarly, the equilibrium constant of the overall reaction is the product of
the equilibrium constants of the two steps.

1 Cu2S(s) → 2 Cu(s) + S(s) ΔG° = + 86.2 kJ ΔH° = + 76.3 kJ

2 S(s) + O2(g)→ SO2(g) ΔG° = –300.1 kJ ΔH° = + 296.8 kJ

3 Cu2S(s)→ 2 Cu(s) + SO2(g) ΔG° = –213.9 kJ ΔH° = – 217.3 kJ

In the above example, reaction 1 is the first step in obtaining metallic copper from one of its principal ores. This reaction is
endothermic and it has a positive Gibbs energy change, so it will not proceed spontaneously at any temperature. If Cu2S is heated
in the air, however, the sulfur is removed as rapidly as it is formed by oxidation in the highly spontaneous reaction 2, which
supplies the Gibbs energy required to drive 1. The combined process, known as roasting, is of considerable industrial importance
and is one of a large class of processes employed for winning metals from their ores.

1.4: Free Energy and Equilibrium (complete) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.
15.6: Free Energy and Equilibrium by Stephen Lower is licensed CC BY 3.0. Original source:
http://www.chem1.com/acad/webtext/virtualtextbook.html.

1.4.7 https://chem.libretexts.org/@go/page/13348
1.5: Electronic Configurations - Spin Quantum Number
The Spin Quantum Number (m ) describes the angular momentum of an electron. An electron spins around an axis and has both
s

angular momentum and orbital angular momentum. Because angular momentum is a vector, the Spin Quantum Number (s) has
both a magnitude (1/2) and direction (+ or -).
Each orbital can only hold two electrons. One electron will have a +1/2 spin and the other will have a -1/2 spin. Electrons like to
fill orbitals before they start to pair up. Therefore the first electron in an orbital will have a spin of +1/2. After all the orbitals are
half filled, the electrons start to pair up. This second electron in the orbital will have a spin of -1/2. If there are two electrons in the
same orbital, it will spin in opposite directions.

Combinations of Quantum Numbers


The three quantum numbers (n, l, and m) that describe an orbital are integers: 0, 1, 2, 3.
The principal quantum number (n) cannot be zero. The allowed values of n are therefore 1, 2, 3, 4...
The angular quantum number (l) can be any integer between 0 and n - 1.
If n = 3, l can be either 0, 1, or 2.
The magnetic quantum number (m) can be any integer between -l and +l.
If l = 2, m can be -2, -1, 0, +1, or +2.
Orbitals that have same value of principal quantum number form a Shell(n).
Orbitals within the shells are divided into subshell (l)
s:l = 0 p:l = 1 d:l = 2 f:l = 3

 Exercise 1.5.1: Tungsten


What is the spin quantum number for Tungsten (symbol W)?

Answer
Tungsten has 4 electrons in the 5d orbital. Therefore 1 electron will go into each orbital (no pairing). The 4th electron will
have a +1/2 spin.

 Exercise 1.5.2: Gold


What is the spin quantum number for Gold (symbol Au)?

Answer
Gold has 9 electrons in the 5d orbital. Therefore the electrons will start to pair up, which means the 9th electron will pair
up, giving it a -1/2 spin.

 Exercise 1.5.3: Sulfur

What is the spin quantum number for Sulfur (symbol S)?

Answer
Sulfur has 4 electrons in the 3p orbitals. The 4th electron in this orbital will be the first one to pair up with another electron,
therefore giving it a -1/2 spin.

References
1. Housecroft, Catherine E., and Alan G. Sharpe. Inorganic Chemistry. 3rd ed. Harlow: Pearson Education, 2008. Print. (pg 15).
2. Nostrand, Van. Encyclopedia of Chemistry. 5th ed. John Wiley and Sons, Inc., 2005. Print. (pg 1396).

1.5.1 https://chem.libretexts.org/@go/page/13619
1.5: Electronic Configurations - Spin Quantum Number is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated
by LibreTexts.
Spin Quantum Number is licensed CC BY 4.0.

1.5.2 https://chem.libretexts.org/@go/page/13619
1.6: Electronic Configurations - Hund's Rule
The Aufbau section discussed how electrons fill the lowest energy orbitals first, and then move up to higher energy orbitals only
after the lower energy orbitals are full. However, there is a problem with this rule. Certainly, 1s orbitals should be filled before 2s
orbitals, because the 1s orbitals have a lower value of n , and thus a lower energy. What about filling the three different 2p orbitals?
In what order should they be filled? The answer to this question involves Hund's rule.
Hund's rule states that:
1. Every orbital in a sublevel is singly occupied before any orbital is doubly occupied.
2. All of the electrons in singly occupied orbitals have the same spin (to maximize total spin).
When assigning electrons to orbitals, an electron first seeks to fill all the orbitals with similar energy (also referred to as degenerate
orbitals) before pairing with another electron in a half-filled orbital. Atoms at ground states tend to have as many unpaired
electrons as possible. In visualizing this process, consider how electrons exhibit the same behavior as the same poles on a magnet
would if they came into contact; as the negatively charged electrons fill orbitals, they first try to get as far as possible from each
other before having to pair up.

 Example 1.6.1: Nitrogen Atoms


Consider the correct electron configuration of the nitrogen (Z = 7) atom: 1s2 2s2 2p3

The p orbitals are half-filled; there are three electrons and three p orbitals. This is because the three electrons in the 2p subshell
will fill all the empty orbitals first before pairing with electrons in them.
Keep in mind that elemental nitrogen is found in nature typically as molecular nitrogen, N , which requires molecular orbitals
2

instead of atomic orbitals as demonstrated above.

 Example 1.6.2: Oxygen Atoms

Next, consider oxygen (Z = 8) atom, the element after nitrogen in the same period; its electron configuration is: 1s2 2s2 2p4

Oxygen has one more electron than nitrogen; as the orbitals are all half-filled, the new electron must pair up. Keep in mind that
elemental oxygen is found in nature typically as molecular oxygen, O , which has molecular orbitals instead of atomic orbitals
2

as demonstrated above.

Hund's Rule Explained


According to the first rule, electrons always enter an empty orbital before they pair up. Electrons are negatively charged and, as a
result, they repel each other. Electrons tend to minimize repulsion by occupying their own orbitals, rather than sharing an orbital
with another electron. Furthermore, quantum-mechanical calculations have shown that the electrons in singly occupied orbitals are
less effectively screened or shielded from the nucleus. Electron shielding is further discussed in the next section.
For the second rule, unpaired electrons in singly occupied orbitals have the same spins. Technically speaking, the first electron in a
sublevel could be either "spin-up" or "spin-down." Once the spin of the first electron in a sublevel is chosen, however, the spins of
all of the other electrons in that sublevel depend on that first spin. To avoid confusion, scientists typically draw the first electron,
and any other unpaired electron, in an orbital as "spin-up."

1.6.1 https://chem.libretexts.org/@go/page/13620
 Example 1.6.3: Carbon and Oxygen

Consider the electron configuration for carbon atoms: 1s22s22p2: The two 2s electrons will occupy the same orbital, whereas
the two 2p electrons will be in different orbital (and aligned the same direction) in accordance with Hund's rule.
Consider also the electron configuration of oxygen. Oxygen has 8 electrons. The electron configuration can be written as
1s22s22p4. To draw the orbital diagram, begin with the following observations: the first two electrons will pair up in the 1s
orbital; the next two electrons will pair up in the 2s orbital. That leaves 4 electrons, which must be placed in the 2p orbitals.
According to Hund’s rule, all orbitals will be singly occupied before any is doubly occupied. Therefore, two p orbital get one
electron and one will have two electrons. Hund's rule also stipulates that all of the unpaired electrons must have the same spin.
In keeping with convention, the unpaired electrons are drawn as "spin-up", which gives (Figure 1).

Purpose of Electron Configurations


When atoms come into contact with one another, it is the outermost electrons of these atoms, or valence shell, that will interact
first. An atom is least stable (and therefore most reactive) when its valence shell is not full. The valence electrons are largely
responsible for an element's chemical behavior. Elements that have the same number of valence electrons often have similar
chemical properties.
Electron configurations can also predict stability. An atom is most stable (and therefore unreactive) when all its orbitals are full.
The most stable configurations are the ones that have full energy levels. These configurations occur in the noble gases. The noble
gases are very stable elements that do not react easily with any other elements. Electron configurations can assist in making
predictions about the ways in which certain elements will react, and the chemical compounds or molecules that different elements
will form.

1.6: Electronic Configurations - Hund's Rule is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.
Hund's Rules is licensed CC BY-SA 4.0.

1.6.2 https://chem.libretexts.org/@go/page/13620
1.7: Electronic Configurations - The Aufbau Process
The Aufbau model lets us take an atom and make predictions about its properties. All we need to know is how many protons it has
(and how many electrons, which is the same as the number of protons for a neutral atom). We can predict the properties of the atom
based on our vague idea of where its electrons are and, more importantly, the energy of those electrons.
How electrons fill in their positions around an atom is called the Aufbau Process (German: "building-up" process). The Aufbau
Process is all about keeping electrons at their lowest possible energy and is the direct result of the Pauli Exclusion Principle. A
corollary of Coulomb's law is that the energy of an electron is affected by attractive and repulsive forces. The closer an electron is
to the nucleus, the lower its energy. The closer an electron is to another electron, the higher its energy.
Of course, a basic principle of thermodynamics is that a system will proceed to the lowest energy possible. That means, if an atom
has only one electron, the electron will have quantum numbers that place it at the lowest possible energy. It will be as close as
possible to the positive nucleus.
If an atom has a second electron, it will also be as close as possible to the nucleus. It could have the same quantum numbers as the
first electron, except for spin. There is a trade-off, of course, because those two electrons will be close enough to repel each other.
However, if it is a choice between that and taking a position much farther from the nucleus, the second electron will go ahead and
pair up. These two electrons are often described as being "in the same orbital" and share their first three quantum numbers, so they
are found somewhere in the same region of space. This first orbital, which has no directional restrictions, is called the 1s orbital.

Figure 1.7.1 : The 1s orbital has no directional preference. Like the yellow tennis ball, it can be found in any direction from the
nucleus.
There is only room for one orbital at this distance from the nucleus. A third electron has to occupy another orbital farther away, the
2s orbital. Again, this is a spherical orbital: the electron can be found in any direction. The 2 in 2s means the Principal Quantum
Number is two (corresponding to the second general energy level). The s is a code for other quantum numbers; it means the
electron can be found in any direction, just like the 1s electrons.
The second energy level is large enough to accommodate additional orbitals, but they are a little farther from the nucleus. These are
called the 2p orbitals. They are regions of space along the x, y and z axes. There are three orbitals of this type, and they are just
called px, py and pz to remind us that they are orthogonal to each other.

Figure 1.7.2 : (left) The 2px orbital has a directional preference. Like the red tennis ball, it can be found only along one axis.
(right). The 2py orbital has a directional preference. Like the red tennis ball, it can be found
only along one axis.
Expressed in a different way, an electron with Principal Quantum Number 2 can have four different combinations of its other
quantum numbers. These combinations are denoted 2s, 2px, 2py and 2pz. The three 2p combinations are a little higher in energy
than the 2s orbital.
A third electron will go into the 2s orbital. It is the lowest in energy. What about a fourth? Does it go into the 2s or a 2p? Once
again the pairing energy is not quite as big as the energy jump up to the 2p orbital. The fourth electron pairs up in the 2s orbital.
A fifth electron goes into one of the 2p orbitals. It does not matter which one. We will say it is the px, arbitrarily. A sixth electron
again could either pair up in the px, or it could go into the py. But the py level is really the same as the px, just in a different
direction. The energy is the same. That means a sixth electron will go into the py rather than pair up in the px, where it would
experience extra repulsion.

1.7.1 https://chem.libretexts.org/@go/page/13621
Note that the p orbitals are often drawn a little differently. For example, p orbitals are usually drawn in a way that shows that they
have phase. Either the two lobes are colored differently to show that they are out of phase with each other, or they are shown with
one lobe shaded and the other left blank.

Figure 1.7.3 : The 2px orbital shown with phase information.


This pattern is generally followed for all of the elements. The tally of how many electrons are found in each orbital is called the
electron configuration. For example, hydrogen has only one electron. Its ground state configuration (that means, assuming the
electron has not been excited to another orbital via addition of energy) is 1s1.
On the other hand, an atom with six electrons, such as the element carbon, has the configuration 1s22s22px12py1.

Periodicity
You may already know that electron configuration is the reason the periodic table works the way it does. Mendelev and others
noticed certain elements had very similar properties, and that is because they have very similar electron configurations. Lithium has
configuration 1s22s1 and its alkali sister, sodium, has configuration 1s22s22px22py22pz23s1. In both cases, the last electron added is
an unpaired s electron. The last electron, or last few electrons, added to an atom generally play a strong role in how the atom
behaves. This "frontier" electron is the one at the outer limits of the atom. If the atom is to interact with anything, the frontier
electron will encounter the thing first. In contrast, the "core" electrons closer to the nucleus are more protected from the outside.
The order of electrons in an atom, from lowest to highest energy, is:
1s
2s
2p
There are some shortcuts we take with electron configurations. We tend to abbreviate "filled shells" (meaning all the possible
orbitals with a given Principal Quantum Number are filled with electrons) and filled "sub-shells" (like 2s or 2p). First of all, in the
case of p orbitals, if all the p orbitals are filled, we might just write 2p6 instead of 2px22py22pz2, because there is only one way all
the orbitals could be filled. However, we would not necessarily write 2p2 instead of 2px12py1, because we may wish to make clear
that the configuration does not involve two electrons in one p orbital at that point, as in 2px22py0.
Also, we dispense with orbital labels entirely to ignore core electrons in a filled shell. For example, instead of writing 1s22s1 for
lithium, we can write [He]2s1. Instead of writing 1s22s22px22py22pz23s1 for sodium, we write [Ne]3s1. The [He] means all the
electrons found in a helium atom, which is a noble gas. The [Ne] means all the electrons found in a neon atom. A noble gas is an
unreactive element with a filled shell: helium, neon, argon, krypton, xenon or radon. The electrons beyond the noble gas shell are
called the valence electrons.
Principal Quantum Number 3 actually allows a third set of orbitals. These are called the d orbitals that are a little like p orbitals, but
they are two-dimensional rather than one-dimensional. A d electron, for example, might extend along the x axis and the y axis, but
not in between the axes.
The d orbitals have five allowed orientations with:
dxy extending along the x and y axes,
dxz extending along the x and z axes,
dyz extending along the y and z axes.
The other two orbitals extend between the axes instead, dx2-y2 rotated 45 degrees away from one of the other d orbitals and extends
between the x and y axes. In the same way, you could imagine an orbital between the x and z and between the y and z axes, but that
would make six different orientations. Quantum mechanical rules do not allow that. As a result, two of the possible combinations
collapse into a mathematical sum, making just one orbital. We call this one the dz2 orbital.
However, just as the 2p orbitals are higher in energy than the 2s orbitals, the 3d orbitals are higher in energy than the 3p orbitals.
The 3d level is very similar in energy to the 4s level. For that reason, calcium's last electrons go into a 4s orbital, not a 3d orbital.
Calcium behaves much like magnesium as a result.
The order of electrons in an atom, from lowest to highest energy, is:

1.7.2 https://chem.libretexts.org/@go/page/13621
1s
2s
2p
3s
3p
4s
3d
4p

Reality is Less Simple


The atomic orbital energies do not properly follow this trend as shown pictorially below. For example, the relative energies of the
3d and 4s orbitals switch between calcium and scandium.

Figure 1.7.4 : Energies of Atomic Orbitals as Functions of Nuclear Charge for Neutral Atoms.

 Exercise 1.7.1

Write electron configurations for the following elements.


1. oxygen, O
2. sulfur, S
3. silicon, Si
4. nitrogen, N
5. argon, Ar
6. neon, Ne

 Exercise 1.7.2

Write abbreviated electron configurations for the following elements.


1. chlorine, Cl
2. calcium, Ca
3. aluminum, Al
4. phosphorus, P

 Exercise 1.7.3
Write abbreviated electron configurations for the following elements.
1. iron, Fe
2. copper, Cu
3. mercury, Hg

1.7.3 https://chem.libretexts.org/@go/page/13621
4. lead, Pb
5. arsenic, As
6. titanium, Ti

Contributors and Attributions


Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

1.7: Electronic Configurations - The Aufbau Process is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.
The Aufbau Process by Chris Schaller is licensed CC BY-NC 3.0.

1.7.4 https://chem.libretexts.org/@go/page/13621
1.8: Electronic Configurations - Pauli Exclusion Principle
The Pauli Exclusion Principle states that, in an atom or molecule, no two electrons can have the same four electronic quantum
numbers. As an orbital can contain a maximum of only two electrons, the two electrons must have opposing spins. This means if
one electron is assigned as a spin up (+1/2) electron, the other electron must be spin-down (-1/2) electron.
Electrons in the same orbital have the same first three quantum numbers, e.g., n = 1 , l = 0 , m = 0 for the 1s subshell. Only two
l

electrons can have these numbers, so that their spin moments must be either m = −1/2 or m = +1/2 . If the 1s orbital contains
s s

only one electron, we have one m value and the electron configuration is written as 1s1 (corresponding to hydrogen). If it is fully
s

occupied, we have two m values, and the electron configuration is 1s2 (corresponding to helium). Visually these two cases can be
s

represented as

As you can see, the 1s and 2s subshells for beryllium atoms can hold only two electrons and when filled, the electrons must have
opposite spins. Otherwise they will have the same four quantum numbers, in violation of the Pauli Exclusion Principle.

1.8: Electronic Configurations - Pauli Exclusion Principle is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated
by LibreTexts.
Pauli Exclusion Principle by Craig Fisher, Sarah Faiz is licensed CC BY 4.0.

1.8.1 https://chem.libretexts.org/@go/page/13622
CHAPTER OVERVIEW
2: Electrochemistry
Electrochemistry is the study of chemical processes that cause electrons to move. This movement of electrons is called electricity,
which can be generated by movements of electrons from one element to another in a reaction known as an oxidation-reduction
("redox") reaction.
2.1: Galvanic Cells
2.2: Standard Electrode Potentials
2.3: Strengths of Oxidants and Reductants
2.4: E, ΔG, and K (The Big Triangle of Chemistry)
2.5: Cell Potential as a Function of Concentrations
2.6: Batteries- Producing Electricity Through Chemical Reactions
2.7: Corrosion- Unwanted Voltaic Cells
2.8: Electrolysis- Causing Nonspontaneous Reactions to Occur

2: Electrochemistry is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

1
2.1: Galvanic Cells
 Learning Objectives
To understand the basics of voltaic cells
To connect voltage from a voltaic cell to underlying redox chemistry

In any electrochemical process, electrons flow from one chemical substance to another, driven by an oxidation–reduction (redox)
reaction. A redox reaction occurs when electrons are transferred from a substance that is oxidized to one that is being reduced. The
reductant is the substance that loses electrons and is oxidized in the process; the oxidant is the species that gains electrons and is
reduced in the process. The associated potential energy is determined by the potential difference between the valence electrons in
atoms of different elements.
Because it is impossible to have a reduction without an oxidation and vice versa, a redox reaction can be described as two half-
reactions, one representing the oxidation process and one the reduction process. For the reaction of zinc with bromine, the overall
chemical reaction is as follows:
2 + −
Zn(s) + Br (aq) → Zn (aq) + 2 Br (aq)
2

The half-reactions are as follows:


reduction half-reaction:
− −
Br (aq) + 2 e → 2 Br (aq)
2

oxidation half-reaction:
2 + −
Zn(s) → Zn (aq) + 2 e

Each half-reaction is written to show what is actually occurring in the system; Zn is the reductant in this reaction (it loses
electrons), and Br is the oxidant (it gains electrons). Adding the two half-reactions gives the overall chemical reaction (Equation
2

2.1.1). A redox reaction is balanced when the number of electrons lost by the reductant equals the number of electrons gained by

the oxidant. Like any balanced chemical equation, the overall process is electrically neutral; that is, the net charge is the same on
both sides of the equation.

In any redox reaction, the number of electrons lost by the oxidation reaction(s) equals the
number of electrons gained by the reduction reaction(s).
In most of our discussions of chemical reactions, we have assumed that the reactants are in intimate physical contact with one
another. Acid–base reactions, for example, are usually carried out with the acid and the base dispersed in a single phase, such as a
liquid solution. With redox reactions, however, it is possible to physically separate the oxidation and reduction half-reactions in
space, as long as there is a complete circuit, including an external electrical connection, such as a wire, between the two half-
reactions. As the reaction progresses, the electrons flow from the reductant to the oxidant over this electrical connection, producing
an electric current that can be used to do work. An apparatus that is used to generate electricity from a spontaneous redox reaction
or, conversely, that uses electricity to drive a nonspontaneous redox reaction is called an electrochemical cell.
There are two types of electrochemical cells: galvanic cells and electrolytic cells. Galvanic cells are named for the Italian physicist
and physician Luigi Galvani (1737–1798), who observed that dissected frog leg muscles twitched when a small electric shock was
applied, demonstrating the electrical nature of nerve impulses. A galvanic (voltaic) cell uses the energy released during a
spontaneous redox reaction (ΔG < 0 ) to generate electricity. This type of electrochemical cell is often called a voltaic cell after its
inventor, the Italian physicist Alessandro Volta (1745–1827). In contrast, an electrolytic cell consumes electrical energy from an
external source, using it to cause a nonspontaneous redox reaction to occur (ΔG > 0 ). Both types contain two electrodes, which
are solid metals connected to an external circuit that provides an electrical connection between the two parts of the system (Figure
2.1.1). The oxidation half-reaction occurs at one electrode (the anode), and the reduction half-reaction occurs at the other (the

cathode). When the circuit is closed, electrons flow from the anode to the cathode. The electrodes are also connected by an
electrolyte, an ionic substance or solution that allows ions to transfer between the electrode compartments, thereby maintaining the
system’s electrical neutrality. In this section, we focus on reactions that occur in galvanic cells.

2.1.1 https://chem.libretexts.org/@go/page/48478
Figure 2.1.1 : Electrochemical Cells. A galvanic cell (left) transforms the energy released by a spontaneous redox reaction into
electrical energy that can be used to perform work. The oxidative and reductive half-reactions usually occur in separate
compartments that are connected by an external electrical circuit; in addition, a second connection that allows ions to flow between
the compartments (shown here as a vertical dashed line to represent a porous barrier) is necessary to maintain electrical neutrality.
The potential difference between the electrodes (voltage) causes electrons to flow from the reductant to the oxidant through the
external circuit, generating an electric current. In an electrolytic cell (right), an external source of electrical energy is used to
generate a potential difference between the electrodes that forces electrons to flow, driving a nonspontaneous redox reaction; only a
single compartment is employed in most applications. In both kinds of electrochemical cells, the anode is the electrode at which the
oxidation half-reaction occurs, and the cathode is the electrode at which the reduction half-reaction occurs.

Voltaic (Galvanic) Cells


To illustrate the basic principles of a galvanic cell, let’s consider the reaction of metallic zinc with cupric ion (Cu2+) to give copper
metal and Zn2+ ion. The balanced chemical equation is as follows:
2 + 2 +
Zn(s) + Cu (aq) → Zn (aq) + Cu(s) (2.1.1)

We can cause this reaction to occur by inserting a zinc rod into an aqueous solution of copper(II) sulfate. As the reaction proceeds,
the zinc rod dissolves, and a mass of metallic copper forms. These changes occur spontaneously, but all the energy released is in the
form of heat rather than in a form that can be used to do work.

I'm not a robot


reCAPTCHA
Privacy - Terms

About this page

Our systems have detected unusual traffic from your


computer network This page checks to see if it's really you
Figure 2.1.2 : The Reaction of Metallic Zinc with Aqueous Copper(II) Ions in a Single Compartment. When a zinc rod is inserted
into a beaker that contains an aqueous solution of copper(II) sulfate, a spontaneous redox reaction occurs: the zinc electrode
dissolves to give Zn (aq) ions, while Cu (aq) ions are simultaneously reduced to metallic copper. The reaction occurs so
2 + 2 +

rapidly that the copper is deposited as very fine particles that appear black, rather than the usual reddish color of copper.
(youtu.be/2gPRK0HmYu4)
This same reaction can be carried out using the galvanic cell illustrated in Figure 2.1.3a. To assemble the cell, a copper strip is
inserted into a beaker that contains a 1 M solution of Cu ions, and a zinc strip is inserted into a different beaker that contains a 1
2 +

M solution of Zn 2 +
ions. The two metal strips, which serve as electrodes, are connected by a wire, and the compartments are
connected by a salt bridge, a U-shaped tube inserted into both solutions that contains a concentrated liquid or gelled electrolyte.
The ions in the salt bridge are selected so that they do not interfere with the electrochemical reaction by being oxidized or reduced
themselves or by forming a precipitate or complex; commonly used cations and anions are Na or K and NO or SO ,+ + −
3
2 −

respectively. (The ions in the salt bridge do not have to be the same as those in the redox couple in either compartment.) When the
circuit is closed, a spontaneous reaction occurs: zinc metal is oxidized to Zn ions at the zinc electrode (the anode), and Cu
2 + 2 +

2.1.2 https://chem.libretexts.org/@go/page/48478
ions are reduced to Cu metal at the copper electrode (the cathode). As the reaction progresses, the zinc strip dissolves, and the
concentration of Zn 2 +
ions in the solution increases; simultaneously, the copper strip gains mass, and the concentration of Cu 2 +

ions in the solution decreases (Figure 2.1.3b). Thus we have carried out the same reaction as we did using a single beaker, but this
time the oxidative and reductive half-reactions are physically separated from each other. The electrons that are released at the anode
flow through the wire, producing an electric current. Galvanic cells therefore transform chemical energy into electrical energy that
can then be used to do work.

Figure 2.1.3 : The Reaction of Metallic Zinc with Aqueous Copper(II) Ions in a Galvanic Cell. (a) A galvanic cell can be
constructed by inserting a copper strip into a beaker that contains an aqueous 1 M solution of Cu2+ ions and a zinc strip into a
different beaker that contains an aqueous 1 M solution of Zn2+ ions. The two metal strips are connected by a wire that allows
electricity to flow, and the beakers are connected by a salt bridge. When the switch is closed to complete the circuit, the zinc
electrode (the anode) is spontaneously oxidized to Zn2+ ions in the left compartment, while Cu2+ ions are simultaneously reduced
to copper metal at the copper electrode (the cathode). (b) As the reaction progresses, the Zn anode loses mass as it dissolves to give
Zn2+(aq) ions, while the Cu cathode gains mass as Cu2+(aq) ions are reduced to copper metal that is deposited on the cathode. (CC
BY-SA-NC; anonymous)
The electrolyte in the salt bridge serves two purposes: it completes the circuit by carrying electrical charge and maintains electrical
neutrality in both solutions by allowing ions to migrate between them. The identity of the salt in a salt bridge is unimportant, as
long as the component ions do not react or undergo a redox reaction under the operating conditions of the cell. Without such a
connection, the total positive charge in the Zn solution would increase as the zinc metal dissolves, and the total positive charge
2 +

in the Cu solution would decrease. The salt bridge allows charges to be neutralized by a flow of anions into the Zn
2 +
solution
2 +

and a flow of cations into the Cu solution. In the absence of a salt bridge or some other similar connection, the reaction would
2 +

rapidly cease because electrical neutrality could not be maintained.


A voltmeter can be used to measure the difference in electrical potential between the two compartments. Opening the switch that
connects the wires to the anode and the cathode prevents a current from flowing, so no chemical reaction occurs. With the switch
closed, however, the external circuit is closed, and an electric current can flow from the anode to the cathode. The potential (E ) cell

of the cell, measured in volts, is the difference in electrical potential between the two half-reactions and is related to the energy
needed to move a charged particle in an electric field. In the cell we have described, the voltmeter indicates a potential of 1.10 V
(Figure 2.1.3a). Because electrons from the oxidation half-reaction are released at the anode, the anode in a galvanic cell is
negatively charged. The cathode, which attracts electrons, is positively charged.
Not all electrodes undergo a chemical transformation during a redox reaction. The electrode can be made from an inert, highly
conducting metal such as platinum to prevent it from reacting during a redox process, where it does not appear in the overall
electrochemical reaction. This phenomenon is illustrated in Example 2.1.1.

A galvanic (voltaic) cell converts the energy released by a spontaneous chemical reaction
to electrical energy. An electrolytic cell consumes electrical energy from an external

2.1.3 https://chem.libretexts.org/@go/page/48478
source to drive a nonspontaneous chemical reaction.

 Example 2.1.1

A chemist has constructed a galvanic cell consisting of two beakers. One beaker contains a strip of tin immersed in aqueous
sulfuric acid, and the other contains a platinum electrode immersed in aqueous nitric acid. The two solutions are connected by
a salt bridge, and the electrodes are connected by a wire. Current begins to flow, and bubbles of a gas appear at the platinum
electrode. The spontaneous redox reaction that occurs is described by the following balanced chemical equation:
− + 2 +
3 Sn(s) + 2 NO (aq) + 8 H (aq) → 3 Sn (aq) + 2 NO(g) + 4 H O(l)
3 2

For this galvanic cell,


a. write the half-reaction that occurs at each electrode.
b. indicate which electrode is the cathode and which is the anode.
c. indicate which electrode is the positive electrode and which is the negative electrode.
Given: galvanic cell and redox reaction
Asked for: half-reactions, identity of anode and cathode, and electrode assignment as positive or negative

Strategy:
A. Identify the oxidation half-reaction and the reduction half-reaction. Then identify the anode and cathode from the half-
reaction that occurs at each electrode.
B. From the direction of electron flow, assign each electrode as either positive or negative.

Solution
A In the reduction half-reaction, nitrate is reduced to nitric oxide. (The nitric oxide would then react with oxygen in the air to
form NO2, with its characteristic red-brown color.) In the oxidation half-reaction, metallic tin is oxidized. The half-reactions
corresponding to the actual reactions that occur in the system are as follows:
reduction:
− + −
NO3 (aq) + 4 H (aq) + 3 e → NO(g) + 2 H O(l)
2

oxidation:
2 + −
Sn(s) → Sn (aq) + 2 e

Thus nitrate is reduced to NO, while the tin electrode is oxidized to Sn2+.
Because the reduction reaction occurs at the Pt electrode, it is the cathode. Conversely, the oxidation reaction occurs at the tin
electrode, so it is the anode.
B Electrons flow from the tin electrode through the wire to the platinum electrode, where they transfer to nitrate. The electric
circuit is completed by the salt bridge, which permits the diffusion of cations toward the cathode and anions toward the anode.
Because electrons flow from the tin electrode, it must be electrically negative. In contrast, electrons flow toward the Pt
electrode, so that electrode must be electrically positive.

 Exercise 2.1.1
Consider a simple galvanic cell consisting of two beakers connected by a salt bridge. One beaker contains a solution of MnO −

in dilute sulfuric acid and has a Pt electrode. The other beaker contains a solution of Sn in dilute sulfuric acid, also with a Pt
2 +

electrode. When the two electrodes are connected by a wire, current flows and a spontaneous reaction occurs that is described
by the following balanced chemical equation:
− 2 + + 2 + 4 +
2 MnO (aq) + 5 Sn (aq) + 16 H (aq) → 2 Mn (aq) + 5 Sn (aq) + 8 H O(l)
4 2

2.1.4 https://chem.libretexts.org/@go/page/48478
For this galvanic cell,
a. write the half-reaction that occurs at each electrode.
b. indicate which electrode is the cathode and which is the anode.
c. indicate which electrode is positive and which is negative.

Answer a
− + − 2 +
MnO (aq) + 8 H (aq) + 5 e → Mn (aq) + 4 H O(l)
4 2

2 + 4 + −
Sn (aq) → Sn (aq) + 2 e

Answer b
The Pt electrode in the permanganate solution is the cathode; the one in the tin solution is the anode.
Answer c
The cathode (electrode in beaker that contains the permanganate solution) is positive, and the anode (electrode in beaker
that contains the tin solution) is negative.

I'm not a robot


reCAPTCHA
Privacy - Terms

About this page

Our systems have detected unusual traffic from your


computer network This page checks to see if it's really you
Electrochemical Cells: Electrochemical Cells(opens in new window) [youtu.be]

Constructing Cell Diagrams (Cell Notation)


Because it is somewhat cumbersome to describe any given galvanic cell in words, a more convenient notation has been developed.
In this line notation, called a cell diagram, the identity of the electrodes and the chemical contents of the compartments are
indicated by their chemical formulas, with the anode written on the far left and the cathode on the far right. Phase boundaries are
shown by single vertical lines, and the salt bridge, which has two phase boundaries, by a double vertical line. Thus the cell diagram
for the Zn/Cu cell shown in Figure 2.1.3a is written as follows:

Figure 2.1.4 : A cell diagram includes solution concentrations when they are provided. The + M term is meant to indicate the
applicable concentration of the species. If the species is a gas, then you substitute the pressure instead.
At the anode is solid zinc. after the phase boundary is aq Zinc two plus and plus M. After the two phase boundary is aq copper two
plus and plus M. At the cathode is solid copper.
Galvanic cells can have arrangements other than the examples we have seen so far. For example, the voltage produced by a redox
reaction can be measured more accurately using two electrodes immersed in a single beaker containing an electrolyte that
completes the circuit. This arrangement reduces errors caused by resistance to the flow of charge at a boundary, called the junction
potential. One example of this type of galvanic cell is as follows:

2.1.5 https://chem.libretexts.org/@go/page/48478
Pt(s) | H (g)|HCl(aq, 1 M) | AgCl(s) Ag(s)
2

This cell diagram does not include a double vertical line representing a salt bridge because there is no salt bridge providing a
junction between two dissimilar solutions. Moreover, solution concentrations have not been specified, so they are not included in
the cell diagram. The half-reactions and the overall reaction for this cell are as follows:
cathode reaction:
− −
AgCl(s) + e → Ag(s) + Cl (aq)

anode reaction:
1 + −
H (g) ⟶ H (aq) + e
2 2

overall:
1 − +
AgCl(s) + H (g) ⟶ Ag(s) + Cl +H (aq)
2 2

A single-compartment galvanic cell will initially exhibit the same voltage as a galvanic cell constructed using separate
compartments, but it will discharge rapidly because of the direct reaction of the reactant at the anode with the oxidized member of
the cathodic redox couple. Consequently, cells of this type are not particularly useful for producing electricity.

 Example 2.1.2

Draw a cell diagram for the galvanic cell described in Example 2.1.1. The balanced chemical reaction is as follows:
− + 2 +
3 Sn(s) + 2 NO (aq) + 8 H (aq) → 3 Sn (aq) + 2 NO(g) + 4 H O(l)
3 2

Given: galvanic cell and redox reaction


Asked for: cell diagram

Strategy:
Using the symbols described, write the cell diagram beginning with the oxidation half-reaction on the left.

Solution
The anode is the tin strip, and the cathode is the Pt electrode. Beginning on the left with the anode, we indicate the phase
boundary between the electrode and the tin solution by a vertical bar. The anode compartment is thus Sn(s) ∣ Sn (aq) . We 2 +

could include H SO (aq) with the contents of the anode compartment, but the sulfate ion (as HSO ) does not participate in
2 4

the overall reaction, so it does not need to be specifically indicated. The cathode compartment contains aqueous nitric acid,
which does participate in the overall reaction, together with the product of the reaction (NO) and the Pt electrode. These are
written as HNO (aq) ∣ NO(g) ∣ Pt(s) , with single vertical bars indicating the phase boundaries. Combining the two
3

compartments and using a double vertical bar to indicate the salt bridge,
2 +
Sn(s) | Sn (aq) || HNO (aq) | NO(g) | Pt(s)
3

The solution concentrations were not specified, so they are not included in this cell diagram.

 Exercise 2.1.2

Draw the cell diagram for the following reaction, assuming the concentration of Ag and Mg + 2 +
are each 1 M:
+ 2 +
Mg(s) + 2 Ag (aq) → Mg (aq) + 2 Ag(s)

Answer
2 + +
Mg(s) | Mg (aq, 1 M) || Ag (aq, 1 M) | Ag(s)

2.1.6 https://chem.libretexts.org/@go/page/48478
I'm not a robot
reCAPTCHA
Privacy - Terms

About this page

Our systems have detected unusual traffic from your


computer network This page checks to see if it's really you
Cell Diagrams: Cell Diagrams(opens in new window) [youtu.be]

Summary
A galvanic (voltaic) cell uses the energy released during a spontaneous redox reaction to generate electricity, whereas an
electrolytic cell consumes electrical energy from an external source to force a reaction to occur. Electrochemistry is the study of the
relationship between electricity and chemical reactions. The oxidation–reduction reaction that occurs during an electrochemical
process consists of two half-reactions, one representing the oxidation process and one the reduction process. The sum of the half-
reactions gives the overall chemical reaction. The overall redox reaction is balanced when the number of electrons lost by the
reductant equals the number of electrons gained by the oxidant. An electric current is produced from the flow of electrons from the
reductant to the oxidant. An electrochemical cell can either generate electricity from a spontaneous redox reaction or consume
electricity to drive a nonspontaneous reaction. In a galvanic (voltaic) cell, the energy from a spontaneous reaction generates
electricity, whereas in an electrolytic cell, electrical energy is consumed to drive a nonspontaneous redox reaction. Both types of
cells use two electrodes that provide an electrical connection between systems that are separated in space. The oxidative half-
reaction occurs at the anode, and the reductive half-reaction occurs at the cathode. A salt bridge connects the separated solutions,
allowing ions to migrate to either solution to ensure the system’s electrical neutrality. A voltmeter is a device that measures the
flow of electric current between two half-reactions. The potential of a cell, measured in volts, is the energy needed to move a
charged particle in an electric field. An electrochemical cell can be described using line notation called a cell diagram, in which
vertical lines indicate phase boundaries and the location of the salt bridge. Resistance to the flow of charge at a boundary is called
the junction potential.

2.1: Galvanic Cells is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
20.3: Voltaic Cells is licensed CC BY-NC-SA 3.0.

2.1.7 https://chem.libretexts.org/@go/page/48478
2.2: Standard Electrode Potentials
Skills to Develop
To use redox potentials to predict whether a reaction is spontaneous.
To balance redox reactions using half-reactions.

In a galvanic cell, current is produced when electrons flow externally through the circuit from the anode to the cathode because of a
difference in potential energy between the two electrodes in the electrochemical cell. In the Zn/Cu system, the valence electrons in
zinc have a substantially higher potential energy than the valence electrons in copper because of shielding of the s electrons of zinc
by the electrons in filled d orbitals. Hence electrons flow spontaneously from zinc to copper(II) ions, forming zinc(II) ions and
metallic copper. Just like water flowing spontaneously downhill, which can be made to do work by forcing a waterwheel, the flow
of electrons from a higher potential energy to a lower one can also be harnessed to perform work.

Figure 2.2.1 : Potential Energy Difference in the Zn/Cu System. The potential energy of a system consisting of metallic Zn and
aqueous Cu2+ ions is greater than the potential energy of a system consisting of metallic Cu and aqueous Zn ions. Much of this
2+

potential energy difference is because the valence electrons of metallic Zn are higher in energy than the valence electrons of
metallic Cu. Because the Zn(s) + Cu2+(aq) system is higher in energy by 1.10 V than the Cu(s) + Zn2+(aq) system, energy is
released when electrons are transferred from Zn to Cu2+ to form Cu and Zn2+.
Because the potential energy of valence electrons differs greatly from one substance to another, the voltage of a galvanic cell
depends partly on the identity of the reacting substances. If we construct a galvanic cell similar to the one in Figure 2.2.1a but
instead of copper use a strip of cobalt metal and 1 M Co2+ in the cathode compartment, the measured voltage is not 1.10 V but 0.51
V. Thus we can conclude that the difference in potential energy between the valence electrons of cobalt and zinc is less than the
difference between the valence electrons of copper and zinc by 0.59 V.
The measured potential of a cell also depends strongly on the concentrations of the reacting species and the temperature of the
system. To develop a scale of relative potentials that will allow us to predict the direction of an electrochemical reaction and the
magnitude of the driving force for the reaction, the potentials for oxidations and reductions of different substances must be
measured under comparable conditions. To do this, chemists use the standard cell potential (E ), defined as the potential of a
o
cell

cell measured under standard conditions—that is, with all species in their standard states (1 M for solutions,Concentrated solutions
of salts (about 1 M) generally do not exhibit ideal behavior, and the actual standard state corresponds to an activity of 1 rather than
a concentration of 1 M. Corrections for nonideal behavior are important for precise quantitative work but not for the more
qualitative approach that we are taking here. 1 atm for gases, pure solids or pure liquids for other substances) and at a fixed
temperature, usually 25°C.

2.2.1 https://chem.libretexts.org/@go/page/48479
Measured redox potentials depend on the potential energy of valence electrons, the
concentrations of the species in the reaction, and the temperature of the system.

Measuring Standard Electrode Potentials


It is physically impossible to measure the potential of a single electrode: only the difference between the potentials of two
electrodes can be measured (this is analogous to measuring absolute enthalpies or free energies since only differences in enthalpy
and free energy can be measured). We can, however, compare the standard cell potentials for two different galvanic cells that have
one kind of electrode in common. This allows us to measure the potential difference between two dissimilar electrodes. For
example, the measured standard cell potential (E ) for the Zn/Cu system is 1.10 V, whereas E for the corresponding Zn/Co
o o

system is 0.51 V. This implies that the potential difference between the Co and Cu electrodes is 1.10 V − 0.51 V = 0.59 V. In fact,
that is exactly the potential measured under standard conditions if a cell is constructed with the following cell diagram:
2 + 2 +
Co(s)| Co (aq, 1 M)|| Cu (aq, 1 M)|Cu(s) E° = 0.59 V (2.2.1)

This cell diagram corresponds to the oxidation of a cobalt anode and the reduction of Cu 2 +
in solution at the copper cathode.
All tabulated values of standard electrode potentials by convention are listed for a reaction written as a reduction, not as an
oxidation, to be able to compare standard potentials for different substances (Table P1). The standard cell potential (E ) is cell
o

therefore the difference between the tabulated reduction potentials of the two half-reactions, not their sum:
E °cell = E °cathode − E °anode (2.2.2)

In contrast, recall that half-reactions are written to show the reduction and oxidation reactions that actually occur in the cell, so the
overall cell reaction is written as the sum of the two half-reactions. According to Equation 2.2.2, when we know the standard
potential for any single half-reaction, we can obtain the value of the standard potential of many other half-reactions by measuring
the standard potential of the corresponding cell.

The overall cell reaction is the sum of the two half-reactions, but the cell potential is the difference between the reduction
potentials:

E °cell = E °cathode − E °anode

Figure 2.2.2 : The Standard Hydrogen Electrode. The SHE consists of platinum wire that is connected to a Pt surface in contact
with an aqueous solution containing 1 M H+ in equilibrium with H2 gas at a pressure of 1 atm. In the molecular view, the Pt surface
catalyzes the oxidation of hydrogen molecules to protons or the reduction of protons to hydrogen gas. (Water is omitted for clarity.)
The standard potential of the SHE is arbitrarily assigned a value of 0 V.
Although it is impossible to measure the potential of any electrode directly, we can choose a reference electrode whose potential is
defined as 0 V under standard conditions. The standard hydrogen electrode (SHE) is universally used for this purpose and is

2.2.2 https://chem.libretexts.org/@go/page/48479
assigned a standard potential of 0 V. It consists of a strip of platinum wire in contact with an aqueous solution containing 1 M H +

(since it is in standard conditions). The H in solution is in equilibrium with H gas at a pressure of 1 atm at the Pt-solution
+
2

interface (Figure 2.2.2). Protons are reduced or hydrogen molecules are oxidized at the Pt surface according to the following
equation:
+ −
2H (aq) + 2 e −
↽⇀
− H (g) (2.2.3)
2

One especially attractive feature of the SHE is that the Pt metal electrode is not consumed during the reaction.

Figure 2.2.3 : Determining a Standard Electrode Potential Using a Standard Hydrogen Electrode. The voltmeter shows that the
standard cell potential of a galvanic cell consisting of a SHE and a Zn/Zn2+ couple is E° = 0.76 V. Because the zinc electrode in
cel l

this cell dissolves spontaneously to form Zn2+(aq) ions while H+(aq) ions are reduced to H2 at the platinum surface, the standard
electrode potential of the Zn2+/Zn couple is −0.76 V.
Figure 2.2.3 shows a galvanic cell that consists of a SHE in one beaker and a Zn strip in another beaker containing a solution of
Zn2+ ions. When the circuit is closed, the voltmeter indicates a potential of 0.76 V. The zinc electrode begins to dissolve to form
Zn2+, and H+ ions are reduced to H2 in the other compartment. Thus the hydrogen electrode is the cathode, and the zinc electrode is
the anode. The diagram for this galvanic cell is as follows:
2 + +
Zn(s)| Zn (aq)|| H (aq, 1 M)| H (g, 1 atm)|Pt(s)
2

The half-reactions that actually occur in the cell and their corresponding electrode potentials are as follows:
cathode:
+ −
2H (aq) + 2 e ⟶ H (g)   E °cathode = 0 V
2

anode:
2 + −
Zn(s) ⟶ Zn (aq) + 2 e   E °anode = −0.76 V

overall:
+ 2 +
Zn(s) + 2 H (aq) ⟶ Zn (aq) + H (g)
2

Then via Equation 2.2.2 we can calculate the standard cell potential

E °cell = E °cathode − E °anode = 0.76 V

2.2.3 https://chem.libretexts.org/@go/page/48479
Caution: Watch the Sign
Although the reaction at the anode is an oxidation, by convention its tabulated E value is reported as a reduction potential.
o

The potential of a half-reaction measured against the SHE under standard conditions is called the standard electrode
potential for that half-reaction. In this example, the standard reduction potential for Zn (aq) + 2 e → Zn(s) is −0.76 V, 2 + −

which means that the standard electrode potential for the reaction that occurs at the anode, the oxidation of Zn to Zn , often 2 +

called the Zn/Zn redox couple, or the Zn/Zn couple, is −(−0.76 V) = 0.76 V.
2 + 2 +

We must therefore subtract E


o
anode
from E
o
cathode
to obtain E °cell = 0.76 V . This is the origin of the negative sign
in Equation 2.2.2.

Because electrical potential is the energy needed to move a charged particle in an electric field, standard electrode potentials for
half-reactions are intensive properties and do not depend on the amount of substance involved. Consequently, E° values are
independent of the stoichiometric coefficients for the half-reaction, and, most important, the coefficients used to produce a balanced
overall reaction do not affect the value of the cell potential.

E° values do NOT depend on the stoichiometric coefficients for a half-reaction, because it


is an intensive property.

Standard Electrode Potentials


To measure the potential of the Cu/Cu2+ couple, we can construct a galvanic cell analogous to the one shown in Figure 2.2.3 but
containing a Cu/Cu2+ couple in the sample compartment instead of Zn/Zn2+. When we close the circuit this time, the measured
potential for the cell is negative (−0.34 V) rather than positive. The negative value of E° indicates that the direction of
cell

2+
spontaneous electron flow is the opposite of that for the Zn/Zn couple. Hence the reactions that occur spontaneously, indicated by
a positive E° , are the reduction of Cu2+ to Cu at the copper electrode. The copper electrode gains mass as the reaction proceeds,
cell

and H2 is oxidized to H+ at the platinum electrode. In this cell, the copper strip is the cathode, and the hydrogen electrode is the
anode. The cell diagram therefore is written with the SHE on the left and the Cu2+/Cu couple on the right:
+ 2 +
Pt(s)| H (g, 1 atm)| H (aq, 1 M)|| Cu (aq, 1 M)|Cu(s)
2

The half-cell reactions and potentials of the spontaneous reaction are as follows:
Cathode:
2 + −
Cu (aq) + 2 e ⟶ Cu(g) E °cathode = 0.34 V

Anode:
+ −
H (g) ⟶ 2 H (aq) + 2 e E °anode = 0 V
2

Overall:
2 + +
H (g) + Cu (aq) ⟶ 2 H (aq) + Cu(s)
2

Then via Equation 2.2.2 we can calculate the standard cell potential

E°cell = E °cathode − E °anode

= 0.34 V

Thus the standard electrode potential for the Cu2+/Cu couple is 0.34 V.

Calculating Standard Cell Potentials


The standard cell potential for a redox reaction (E°cell) is a measure of the tendency of reactants in their standard states to form
products in their standard states; consequently, it is a measure of the driving force for the reaction, which earlier we called voltage.
We can use the two standard electrode potentials we found earlier to calculate the standard potential for the Zn/Cu cell represented
by the following cell diagram:
2 + 2 +
Zn(s)| Zn (aq, 1 M)|| Cu (aq, 1 M)|Cu(s)

2.2.4 https://chem.libretexts.org/@go/page/48479
We know the values of E°anode for the reduction of Zn2+ and E°cathode for the reduction of Cu2+, so we can calculate E° cell :
cathode:
2 + −
Cu (aq) + 2 e ⟶ Cu(s) E °cathode = 0.34 V

anode:
2 + −
Zn(s) ⟶ Zn (aq, 1 M) + 2 e E °anode = −0.76 V

overall:
2 + 2 +
Zn(s) + Cu (aq) ⟶ Zn (aq) + Cu(s)

Then via Equation 2.2.2 we can calculate the standard cell potential
E°cell = E °cathode − E °anode

= 1.10 V

This is the same value that is observed experimentally. If the value of E° is positive, the reaction will occur spontaneously as
cell

written. If the value of E° cell is negative, then the reaction is not spontaneous, and it will not occur as written under standard
conditions; it will, however, proceed spontaneously in the opposite direction. As we shall see in Section 20.9, this does not mean
that the reaction cannot be made to occur at all under standard conditions. With a sufficient input of electrical energy, virtually any
reaction can be forced to occur. Example 2.2.2 and its corresponding exercise illustrate how we can use measured cell potentials to
calculate standard potentials for redox couples.

A positive E° means that the reaction will occur spontaneously as written. A negative
cell

E° cellmeans that the reaction will proceed spontaneously in the opposite direction.
Example 2.2.1

A galvanic cell with a measured standard cell potential of 0.27 V is constructed using two beakers connected by a salt bridge.
One beaker contains a strip of gallium metal immersed in a 1 M solution of GaCl3, and the other contains a piece of nickel
immersed in a 1 M solution of NiCl . The half-reactions that occur when the compartments are connected are as follows:
2

cathode:
2 + −
Ni (aq) + 2 e ⟶ Ni(s)

anode:
3 + −
Ga(s) → Ga (aq) + 3 e

If the potential for the oxidation of Ga to Ga 3 +


is 0.55 V under standard conditions, what is the potential for the oxidation of
Ni to Ni2+?
Given: galvanic cell, half-reactions, standard cell potential, and potential for the oxidation half-reaction under standard
conditions
Asked for: standard electrode potential of reaction occurring at the cathode
Strategy:
A. Write the equation for the half-reaction that occurs at the anode along with the value of the standard electrode potential for
the half-reaction.
B. Use Equation 2.2.2 to calculate the standard electrode potential for the half-reaction that occurs at the cathode. Then
reverse the sign to obtain the potential for the corresponding oxidation half-reaction under standard conditions.
Solution:
A We have been given the potential for the oxidation of Ga to Ga3+ under standard conditions, but to report the standard
electrode potential, we must reverse the sign. For the reduction reaction Ga3+(aq) + 3e− → Ga(s), E°anode = −0.55 V.

2.2.5 https://chem.libretexts.org/@go/page/48479
B Using the value given for E° and the calculated value of E°anode, we can calculate the standard potential for the reduction
cell

of Ni2+ to Ni from Equation 2.2.2:


E°cell = E °cathode − E °anode

o
0.27 V =E − (−0.55V )
cathode


E = −0.28 V
cathode

This is the standard electrode potential for the reaction Ni2+(aq) + 2e− → Ni(s). Because we are asked for the potential for the
oxidation of Ni to Ni2+ under standard conditions, we must reverse the sign of E°cathode. Thus E° = −(−0.28 V) = 0.28 V for the
oxidation. With three electrons consumed in the reduction and two produced in the oxidation, the overall reaction is not
balanced. Recall, however, that standard potentials are independent of stoichiometry.

Exercise 2.2.1

A galvanic cell is constructed with one compartment that contains a mercury electrode immersed in a 1 M aqueous solution of
mercuric acetate Hg(CH CO ) and one compartment that contains a strip of magnesium immersed in a 1 M aqueous
3 2 2

solution of MgCl . When the compartments are connected, a potential of 3.22 V is measured and the following half-reactions
2

occur:
cathode:
2 + −
Hg (aq) + 2 e ⟶ Hg(l)

anode:
2 + −
Mg(s) ⟶ Mg (aq) + 2 e

If the potential for the oxidation of Mg to Mg 2 +


is 2.37 V under standard conditions, what is the standard electrode potential
for the reaction that occurs at the cathode?

Answer
0.85 V

Reference Electrodes
When using a galvanic cell to measure the concentration of a substance, we are generally interested in the potential of only one of
the electrodes of the cell, the so-called indicator electrode, whose potential is related to the concentration of the substance being
measured. To ensure that any change in the measured potential of the cell is due to only the substance being analyzed, the potential
of the other electrode, the reference electrode, must be constant. You are already familiar with one example of a reference
electrode: the SHE. The potential of a reference electrode must be unaffected by the properties of the solution, and if possible, it
should be physically isolated from the solution of interest. To measure the potential of a solution, we select a reference electrode
and an appropriate indicator electrode. Whether reduction or oxidation of the substance being analyzed occurs depends on the
potential of the half-reaction for the substance of interest (the sample) and the potential of the reference electrode.

The potential of any reference electrode should not be affected by the properties of the
solution to be analyzed, and it should also be physically isolated.
There are many possible choices of reference electrode other than the SHE in Figure 2.2.2. The SHE requires a constant flow of
highly flammable hydrogen gas, which makes it inconvenient to use. Consequently, two other electrodes are commonly chosen as
reference electrodes. One is the silver–silver chloride electrode, which consists of a silver wire coated with a very thin layer of
AgCl that is dipped into a chloride ion solution with a fixed concentration. The silver–silver chloride electrode cell diagram is


Cl (aq)|AgCl(s)|Ag(s)

and corresponding half-reaction is


− −
AgCl(s) + e ⟶ Ag(s) + Cl (aq).

2.2.6 https://chem.libretexts.org/@go/page/48479
If a saturated solution of KCl is used as the chloride solution, the potential of the silver–silver chloride electrode is 0.197 V versus
the SHE. That is, 0.197 V must be subtracted from the measured value to obtain the standard electrode potential measured against
the SHE.
A second common reference electrode is the saturated calomel electrode (SCE), which has the same general form as the silver–
silver chloride electrode. The SCE consists of a platinum wire inserted into a moist paste of liquid mercury (Hg Cl ; called 2 2

calomel in the old chemical literature) and KCl. This interior cell is surrounded by an aqueous KCl solution, which acts as a salt
bridge between the interior cell and the exterior solution (Figure 2.2.5a). Although it sounds and looks complex, this cell is
actually easy to prepare and maintain, and its potential is highly reproducible. The SCE cell diagram is

Pt(s)| Hg Cl (s)|KCl(aq, sat)


2 2

and corresponding half-reaction is


− −
Hg Cl (s) + 2 e ⟶ 2 Hg(l) + 2 Cl (aq).
2 2

At 25°C, the potential of the SCE is 0.2415 V versus the SHE, which means that 0.2415 V must be subtracted from the potential
versus an SCE to obtain the standard electrode potential.

Figure 2.2.5 : Three Common Types of Electrodes. (a) The SCE is a reference electrode that consists of a platinum wire inserted
into a moist paste of liquid mercury (calomel; Hg2Cl2) and KCl. The interior cell is surrounded by an aqueous KCl solution, which
acts as a salt bridge between the interior cell and the exterior solution. (b) In a glass electrode, an internal Ag/AgCl electrode is
immersed in a 1 M HCl solution that is separated from the sample solution by a very thin glass membrane. The potential of the
electrode depends on the H+ ion concentration of the sample. (c) The potential of an ion-selective electrode depends on the
concentration of only a single ionic species in solution.

Using Electrochemistry to Measure Concentrations


One of the most common uses of electrochemistry is to measure the H ion concentration of a solution. A glass electrode is
+

generally used for this purpose, in which an internal Ag/AgCl electrode is immersed in a 0.10 M HCl solution that is separated
from the solution by a very thin glass membrane (Figure 2.2.5b. The glass membrane absorbs protons, which affects the measured
potential. The extent of the adsorption on the inner side is fixed because [H ] is fixed inside the electrode, but the adsorption of
+

protons on the outer surface depends on the pH of the solution. The potential of the glass electrode depends on [H ] as follows +

(recall that pH = − log[H ]):


+

+
Eglass = E' + (0.0591 V × log[ H ])

= E' − 0.0591 V × pH

The voltage E' is a constant that depends on the exact construction of the electrode. Although it can be measured, in practice, a
glass electrode is calibrated; that is, it is inserted into a solution of known pH, and the display on the pH meter is adjusted to the
known value (Figure 2.2.7). Once the electrode is properly calibrated, it can be placed in a solution and used to determine an
unknown pH.

2.2.7 https://chem.libretexts.org/@go/page/48479
Figure 2.2.7 : A pH meter measures the hydrogen-ion concentration in water-based solutions, indicating its acidity or alkalinity
expressed as pH. The pH meter measures the difference in electrical potential between a pH electrode and a reference electrode,
and so the pH meter is sometimes referred to as a "potentiometric pH meter". (CC By-SA 2.0 Generic; Sergei Golyshev via
Wikipedia)
Ion-selective electrodes are used to measure the concentration of a particular species in solution; they are designed so that their
potential depends on only the concentration of the desired species (Figure 2.2.5c). These electrodes usually contain an internal
reference electrode that is connected by a solution of an electrolyte to a crystalline inorganic material or a membrane, which acts as
the sensor. For example, one type of ion-selective electrode uses a single crystal of Eu-doped LaF as the inorganic material. When
3

fluoride ions in solution diffuse to the surface of the solid, the potential of the electrode changes, resulting in a so-called fluoride
electrode. Similar electrodes are used to measure the concentrations of other species in solution. Some of the species whose
concentrations can be determined in aqueous solution using ion-selective electrodes and similar devices are listed in Table 2.2.1.
Table 2.2.1 : Some Species Whose Aqueous Concentrations Can Be Measured Using Electrochemical Methods
Species Type of Sample

H+ laboratory samples, blood, soil, and ground and surface water

NH3/NH4+ wastewater and runoff water

K+ blood, wine, and soil

CO2/HCO3− blood and groundwater

F− groundwater, drinking water, and soil

Br− grains and plant extracts

I− milk and pharmaceuticals

NO3− groundwater, drinking water, soil, and fertilizer

Summary
Redox reactions can be balanced using the half-reaction method. The standard cell potential is a measure of the driving force for
the reaction.
E °cell = E °cathode − E °anode

The flow of electrons in an electrochemical cell depends on the identity of the reacting substances, the difference in the potential
energy of their valence electrons, and their concentrations. The potential of the cell under standard conditions (1 M for solutions, 1
atm for gases, pure solids or liquids for other substances) and at a fixed temperature (25°C) is called the standard cell potential
(E°cell). Only the difference between the potentials of two electrodes can be measured. By convention, all tabulated values of
standard electrode potentials are listed as standard reduction potentials. The overall cell potential is the reduction potential of the
reductive half-reaction minus the reduction potential of the oxidative half-reaction (E°cell = E°cathode − E°anode). The potential of the
standard hydrogen electrode (SHE) is defined as 0 V under standard conditions. The potential of a half-reaction measured against
the SHE under standard conditions is called its standard electrode potential. The standard cell potential is a measure of the driving
force for a given redox reaction. All E° values are independent of the stoichiometric coefficients for the half-reaction. Redox

2.2.8 https://chem.libretexts.org/@go/page/48479
reactions can be balanced using the half-reaction method, in which the overall redox reaction is divided into an oxidation half-
reaction and a reduction half-reaction, each balanced for mass and charge. The half-reactions selected from tabulated lists must
exactly reflect reaction conditions. In an alternative method, the atoms in each half-reaction are balanced, and then the charges are
balanced. Whenever a half-reaction is reversed, the sign of E° corresponding to that reaction must also be reversed. If E° is cell

positive, the reaction will occur spontaneously under standard conditions. If E° is negative, then the reaction is not spontaneous
cell

under standard conditions, although it will proceed spontaneously in the opposite direction. The potential of an indicator electrode
is related to the concentration of the substance being measured, whereas the potential of the reference electrode is held constant.
Whether reduction or oxidation occurs depends on the potential of the sample versus the potential of the reference electrode. In
addition to the SHE, other reference electrodes are the silver–silver chloride electrode; the saturated calomel electrode (SCE); the
glass electrode, which is commonly used to measure pH; and ion-selective electrodes, which depend on the concentration of a
single ionic species in solution. Differences in potential between the SHE and other reference electrodes must be included when
calculating values for E°.

2.2: Standard Electrode Potentials is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

2.2.9 https://chem.libretexts.org/@go/page/48479
2.3: Strengths of Oxidants and Reductants
Learning Objectives
Identify how to view Standard Reduction Potentials from the perspective of viable reducing and oxidizing agents in
REDOX reactions.

We can measure the standard potentials for a wide variety of chemical substances, some of which are listed in Table P2. These data
allow us to compare the oxidative and reductive strengths of a variety of substances. The half-reaction for the standard hydrogen
electrode (SHE) lies more than halfway down the list in Table 2.3.1. All reactants that lie below the SHE in the table are stronger
oxidants than H+, and all those that lie above the SHE are weaker. The strongest oxidant in the table is F2, with a standard electrode
potential of 2.87 V. This high value is consistent with the high electronegativity of fluorine and tells us that fluorine has a stronger
tendency to accept electrons (it is a stronger oxidant) than any other element.
Table 2.3.1 : Standard Potentials for Selected Reduction Half-Reactions at 25°C
Half-Reaction E° (V)

Li+(aq) + e− ⇌ Li(s) –3.040

Be2+(aq) + 2e− ⇌ Be(s) –1.99

Al3+(aq) + 3e− ⇌ Al(s) –1.676

Zn2+(aq) + 2e− ⇌ Zn(s) –0.7618

Ag2S(s) + 2e− ⇌ 2Ag(s) + S2−(aq) –0.71

Fe2+(aq) + 2e− ⇌ Fe(s) –0.44

Cr3+(aq) + e− ⇌ Cr2+(aq) –0.424

Cd2+(aq) + 2e− ⇌ Cd(s) –0.4030

PbSO4(s) + 2e− ⇌ Pb(s) + SO42−(aq) –0.356

Ni2+(aq) + 2e− ⇌ Ni(s) –0.257

2SO42−(aq) + 4H+(aq) + 2e− ⇌ S2O62−(aq) + 2H2O(l) –0.25

Sn2+(aq) + 2e− ⇌ Sn(s) −0.14

2H+(aq) + 2e− ⇌ H2(g) 0.00

Sn4+(aq) + 2e− ⇌ Sn2+(aq) 0.154

Cu2+(aq) + e− ⇌ Cu+(aq) 0.159

AgCl(s) + e− ⇌ Ag(s) + Cl−(aq) 0.2223

Cu2+(aq) + 2e− ⇌ Cu(s) 0.3419

O2(g) + 2H2O(l) + 4e− ⇌ 4OH−(aq) 0.401

H2SO3(aq) + 4H+(aq) + 4e− ⇌ S(s) + 3H2O(l) 0.45

I2(s) + 2e− ⇌ 2I−(aq) 0.5355

MnO42−(aq) + 2H2O(l) + 2e− ⇌ MnO2(s) + 4OH−(aq) 0.60

O2(g) + 2H+(aq) + 2e− ⇌ H2O2(aq) 0.695

H2SeO3(aq) + 4H+ + 4e− ⇌ Se(s) + 3H2O(l) 0.74

Fe3+(aq) + e− ⇌ Fe2+(aq) 0.771

Ag+(aq) + e− ⇌ Ag(s) 0.7996

NO3−(aq) + 3H+(aq) + 2e− ⇌ HNO2(aq) + H2O(l) 0.94

Br2(aq) + 2e− ⇌ 2Br−(aq) 1.087

2.3.1 https://chem.libretexts.org/@go/page/83585
Half-Reaction E° (V)

MnO2(s) + 4H+(aq) + 2e− ⇌ Mn2+(aq) + 2H2O(l) 1.23

O2(g) + 4H+(aq) + 4e− ⇌ 2H2O(l) 1.229

Cr2O72−(aq) + 14H+(aq) + 6e− ⇌ 2Cr3+(aq) + 7H2O(l) 1.36

Cl2(g) + 2e− ⇌ 2Cl−(aq) 1.396

Ce
4+
(aq) + e

⇌ Ce
3+
(aq) 1.61

PbO2(s) + HSO4−(aq) + 3H+(aq) + 2e− ⇌ PbSO4(s) + 2H2O(l) 1.690

H2O2(aq) + 2H+(aq) + 2e− ⇌ 2H2O(l) 1.763

F2(g) + 2e−⇌ 2F−(aq) 2.87

Not all oxidizers and reducers are created equal. The standard reduction potentials in Table 2.3.1 can be interpreted as a ranking of
substances according to their oxidizing and reducing power. Strong oxidizing agents are typically compounds with elements in high
oxidation states or with high electronegativity, which gain electrons in the redox reaction (Figure 2.3.1). Examples of strong
oxidizers include hydrogen peroxide, permanganate, and osmium tetroxide. Reducing agents are typically electropositive elements
such as hydrogen, lithium, sodium, iron, and aluminum, which lose electrons in redox reactions. Hydrides (compounds that contain
hydrogen in the formal -1 oxidation state), such as sodium hydride, sodium borohydride and lithium aluminum hydride, are often
used as reducing agents in organic and organometallic reactions.

Figure 2.3.1 : Table of standard electrode potentials. Note that the ranking is the opposite here, with strong oxidizing agents at
the top and strong reducing agents at the bottom, than in Table 2.3.1 . The actual standard potentials are the same of course. (CC
BY-SA 3.0; Tem5psu).
Similarly, all species in Table 2.3.1 that lie above H2 are stronger reductants than H2, and those that lie below H2 are weaker. The
strongest reductant in the table is thus metallic lithium, with a standard electrode potential of −3.04 V. This fact might be surprising
because cesium, not lithium, is the least electronegative element. The apparent anomaly can be explained by the fact that electrode
potentials are measured in aqueous solution, where intermolecular interactions are important, whereas ionization potentials and
electron affinities are measured in the gas phase. Due to its small size, the Li+ ion is stabilized in aqueous solution by strong
electrostatic interactions with the negative dipole end of water molecules. These interactions result in a significantly greater
ΔHhydration for Li+ compared with Cs+. Lithium metal is therefore the strongest reductant (most easily oxidized) of the alkali metals
in aqueous solution.

The standard reduction potentials can be interpreted as a ranking of substances according


to their oxidizing and reducing power. Species in Table 2.3.1 that lie above H2 are

2.3.2 https://chem.libretexts.org/@go/page/83585
stronger reducing agents (more easily oxidized) than H2. Species that lie below H
2
are
stronger oxidizing agents.
Because the half-reactions shown in Table 2.3.1 are arranged in order of their E° values, we can use the table to quickly predict the
relative strengths of various oxidants and reductants. Any species on the left side of a half-reaction will spontaneously oxidize any
species on the right side of another half-reaction that lies below it in the table. Conversely, any species on the right side of a half-
reaction will spontaneously reduce any species on the left side of another half-reaction that lies above it in the table. We can use
these generalizations to predict the spontaneity of a wide variety of redox reactions (E°cell > 0), as illustrated in Example 2.3.1.

Example 2.3.1: Silver Sulfide

The black tarnish that forms on silver objects is primarily Ag S


2
. The half-reaction for reversing the tarnishing process is as
follows:
− 2 −
Ag S(s) + 2 e → 2 Ag(s) + S (aq)   E° = −0.69 V
2

a. Referring to Table 2.3.1, predict which species—H2O2(aq), Zn(s), I−(aq), Sn2+(aq)—can reduce Ag2S to Ag under standard
conditions.
b. Of these species—H2O2(aq), Zn(s), I−(aq), Sn2+(aq), identify which is the strongest reducing agent in aqueous solution and
thus the best candidate for a commercial product.
c. From the data in Table 2.3.1, suggest an alternative reducing agent that is readily available, inexpensive, and possibly more
effective at removing tarnish.

Given: reduction half-reaction, standard electrode potential, and list of possible reductants
Asked for: reductants for Ag 2
S , strongest reductant, and potential reducing agent for removing tarnish
Strategy:
A From their positions in Table 2.3.1 , decide which species can reduce Ag S
2
. Determine which species is the strongest
reductant.
B Use Table 2.3.1 to identify a reductant for Ag 2
S that is a common household product.
Solution
We can solve the problem in one of two ways: (1) compare the relative positions of the four possible reductants with that of the
Ag2S/Ag couple in Table 2.3.1 or (2) compare E° for each species with E° for the Ag2S/Ag couple (−0.69 V).
a. A The species in Table 2.3.1 are arranged from top to bottom in order of increasing reducing strength. Of the four species
given in the problem, I−(aq), Sn2+(aq), and H2O2(aq) lie above Ag2S, and one [Zn(s)] lies below it. We can therefore
conclude that Zn(s) can reduce Ag2S(s) under standard conditions, whereas I−(aq), Sn2+(aq), and H2O2(aq) cannot. Sn2+(aq)
and H2O2(aq) appear twice in the table: on the left side (oxidant) in one half-reaction and on the right side (reductant) in
another.

2.3.3 https://chem.libretexts.org/@go/page/83585
b. The strongest reductant is Zn(s), the species on the right side of the half-reaction that lies closer to the bottom of Table
− 2+
2.3.1 than the half-reactions involving I (aq), Sn (aq), and H2O2(aq). (Commercial products that use a piece of zinc are
often marketed as a “miracle product” for removing tarnish from silver. All that is required is to add warm water and salt
for electrical conductivity.)
c. B Of the reductants that lie below Zn(s) in Table 2.3.1, and therefore are stronger reductants, only one is commonly
available in household products: Al(s), which is sold as aluminum foil for wrapping foods.

Example 2.3.2

Use the data in Table 2.3.1 to determine whether each reaction is likely to occur spontaneously under standard conditions:
a. Sn(s) + Be (aq) → Sn (aq) + Be(s)
2 + 2 +

b. MnO (s) + H O (aq) + 2 H (aq) → O


2 2 2
+

2
(g) + Mn
2 +
(aq) + 2 H O(l)
2

Given: redox reaction and list of standard electrode potentials (Table 2.3.1)
Asked for: reaction spontaneity
Strategy:
A. Identify the half-reactions in each equation. Using Table 2.3.1, determine the standard potentials for the half-reactions in
the appropriate direction.
B. Use the E = E (cathode) − E (anode) equation to calculate the standard cell potential for the overall reaction.
cell
o
srp
o
srp

From this value, determine whether the overall reaction is spontaneous.


Solution
a. A Metallic tin is oxidized to Sn (aq), and Be (aq) is reduced to elemental beryllium. We can find the standard
2 + 2 +

electrode potentials for the latter (reduction) half-reaction (−1.99 V) and for the former (oxidation) half-reaction (−0.14 V)
directly from Table 2.3.1.
B Adding the two half-reactions gives the overall reaction:
2 + −
cathode : Be (aq) + 2 e → Be(s)

2 + −
anode : Sn(s) → Sn (s) + 2 e

2 + 2 +
overall : Sn(s) + Be (aq) → Sn (aq) + Be(s)

with

E = -1.99 V
cathode


E = -0.14 V
anode

∘ ∘ ∘
E =E −E
cell cathode anode

= −1.85 V

The standard cell potential is quite negative, so the reaction will not occur spontaneously as written. That is, metallic tin
cannot reduce Be2+ to beryllium metal under standard conditions. Instead, the reverse process, the reduction of stannous
ions (Sn2+) by metallic beryllium, which has a positive value of E°cell, will occur spontaneously.
b. A MnO is the oxidant (Mn is reduced to Mn ), while H O is the reductant (O is oxidized to O ). We can obtain
2
4 + 2 +
2 2
2 −

the standard electrode potentials for the reduction and oxidation half-reactions directly from Table 2.3.1.
B The two half-reactions and their corresponding potentials are as follows:
+ − 2 +
cathode : MnO (s) + 4 H (aq) + 2 e → Mn (aq) + 2 H O(l)
2 2
+ −
anode : H O (aq) → O (g) + 2 H (aq) + 2 e
2 2 2

+ 2 +
overall : MnO (s) + H O (aq) + 2 H (aq) → O (g) + Mn (aq) + 2 H O(l)
2 2 2 2 2

with

2.3.4 https://chem.libretexts.org/@go/page/83585

E = 1.23 V
cathode


E = 0.70 V
anode

∘ ∘ ∘
E =E −E
cell cathode anode

= +0.52 V

The standard potential for the reaction is positive, indicating that under standard conditions, it will occur spontaneously as
written. Hydrogen peroxide will reduce MnO , and oxygen gas will evolve from the solution.
2

Exercise 2.3.2

Use the data in Table 2.3.1 to determine whether each reaction is likely to occur spontaneously under standard conditions:
a. 2 Ce (aq) + 2 Cl (aq) → 2 Ce (aq) + Cl (g)
4 + − 3 +

b. 4 MnO (s) + 3 O (g) + 4 OH (aq) → 4 MnO (aq) + 2 H


2 2
− −

4 2
O

Answer a
spontaneous with E o
cell
= 1.61 V − 1.396 V = 0.214 V

Answer b
nonspontaneous with E ∘
cell
= −0.20 V

Although the sign of E o


cell
tells us whether a particular redox reaction will occur spontaneously under standard conditions, it does
not tell us to what extent the reaction proceeds, and it does not tell us what will happen under nonstandard conditions. To answer
these questions requires a more quantitative understanding of the relationship between electrochemical cell potential and chemical
thermodynamics.

Summary
The relative strengths of various oxidants and reductants can be predicted using E values. The oxidative and reductive strengths
o

of a variety of substances can be compared using standard electrode potentials. Apparent anomalies can be explained by the fact
that electrode potentials are measured in aqueous solution, which allows for strong intermolecular electrostatic interactions, and not
in the gas phase.

Conceptual Problems
1. The order of electrode potentials cannot always be predicted by ionization potentials and electron affinities. Why? Do you
expect sodium metal to have a higher or a lower electrode potential than predicted from its ionization potential? What is its
approximate electrode potential?
2. Without referring to tabulated data, of Br2/Br−, Ca2+/Ca, O2/OH−, and Al3+/Al, which would you expect to have the least
negative electrode potential and which the most negative? Why?
3. Because of the sulfur-containing amino acids present in egg whites, eating eggs with a silver fork will tarnish the fork. As a
chemist, you have all kinds of interesting cleaning products in your cabinet, including a 1 M solution of oxalic acid (H2C2O4).
Would you choose this solution to clean the fork that you have tarnished from eating scrambled eggs?
4. The electrode potential for the reaction Cu2+(aq) + 2e− → Cu(s) is 0.34 V under standard conditions. Is the potential for the
oxidation of 0.5 mol of Cu equal to −0.34/2 V? Explain your answer.
5. Refer to Table 2.3.1 to predict
A. Which species—Sn4+(aq), Cl−(aq), Ag+(aq), Cr3+(aq), and/or H2O2(aq)—can oxidize MnO2(s) to MNO4− under standard
conditions.
B. Which species—Sn4+(aq), Cl−(aq), Ag+(aq), Cr3+(aq), and/or H2O2(aq)—is the strongest oxidizing agent in aqueous
solution.

Conceptual Answer
3. No; E° = −0.691 V for Ag2S(s) + 2e− → Ag(s) + S2−(aq), which is too negative for Ag2S to be spontaneously reduced by oxalic
acid [E° = 0.49 V for 2CO2(g) + 2H+(aq) + 2e− → H2C2O4(aq)]

2.3.5 https://chem.libretexts.org/@go/page/83585
5.
A. Ag+(aq); H2O2(aq)
B. H2O2(aq)

2.3: Strengths of Oxidants and Reductants is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
Comparing Strengths of Oxidants and Reductants is licensed CC BY-NC-SA 4.0.

2.3.6 https://chem.libretexts.org/@go/page/83585
2.4: E, ΔG, and K (The Big Triangle of Chemistry)
 Learning Objectives
To understand the relationship between cell potential and the equilibrium constant.
To use cell potentials to calculate solution concentrations.

Changes in reaction conditions can have a tremendous effect on the course of a redox reaction. For example, under standard conditions, the reaction of Co(s) with Ni (aq) to form Ni(s) and 2 +

(aq) occurs spontaneously, but if we reduce the concentration of Ni by a factor of 100, so that [Ni ] is 0.01 M, then the reverse reaction occurs spontaneously instead. The relationship
2 + 2 + 2 +
Co

between voltage and concentration is one of the factors that must be understood to predict whether a reaction will be spontaneous.

The Relationship between Cell Potential & Gibbs Energy


Electrochemical cells convert chemical energy to electrical energy and vice versa. The total amount of energy produced by an electrochemical cell, and thus the amount of energy available to do
electrical work, depends on both the cell potential and the total number of electrons that are transferred from the reductant to the oxidant during the course of a reaction. The resulting electric current
is measured in coulombs (C), an SI unit that measures the number of electrons passing a given point in 1 s. A coulomb relates energy (in joules) to electrical potential (in volts). Electric current is
measured in amperes (A); 1 A is defined as the flow of 1 C/s past a given point (1 C = 1 A·s):
1 J
= 1 C = A⋅s (2.4.1)
1 V

In chemical reactions, however, we need to relate the coulomb to the charge on a mole of electrons. Multiplying the charge on the electron by Avogadro’s number gives us the charge on 1 mol of
electrons, which is called the faraday (F), named after the English physicist and chemist Michael Faraday (1791–1867):
23
−19
6.02214 × 10 J
F = (1.60218 × 10  C) ( ) (2.4.2)

1 mol e

4 −
= 9.64833212 × 10  C/mol e (2.4.3)


≃ 96, 485 J/(V ⋅ mol e ) (2.4.4)

The total charge transferred from the reductant to the oxidant is therefore nF , where n is the number of moles of electrons.

 Michael Faraday (1791–1867)

Faraday was a British physicist and chemist who was arguably one of the greatest experimental scientists in history. The son of a blacksmith, Faraday was self-educated and became an apprentice
bookbinder at age 14 before turning to science. His experiments in electricity and magnetism made electricity a routine tool in science and led to both the electric motor and the electric generator.
He discovered the phenomenon of electrolysis and laid the foundations of electrochemistry. In fact, most of the specialized terms introduced in this chapter (electrode, anode, cathode, and so
forth) are due to Faraday. In addition, he discovered benzene and invented the system of oxidation state numbers that we use today. Faraday is probably best known for “The Chemical History of
a Candle,” a series of public lectures on the chemistry and physics of flames.

The maximum amount of work that can be produced by an electrochemical cell (w max ) is equal to the product of the cell potential (E ∘
cell
) and the total charge transferred during the reaction (nF ):
wmax = nF Ecell (2.4.5)

Work is expressed as a negative number because work is being done by a system (an electrochemical cell with a positive potential) on its surroundings.
The change in free energy (ΔG) is also a measure of the maximum amount of work that can be performed during a chemical process (ΔG = w max ). Consequently, there must be a relationship
between the potential of an electrochemical cell and ΔG; this relationship is as follows:
ΔG = −nF Ecell (2.4.6)

A spontaneous redox reaction is therefore characterized by a negative value of ΔG and a positive value of E ∘
cell
, consistent with our earlier discussions. When both reactants and products are in their
standard states, the relationship between ΔG° and E is as follows:

cell

∘ ∘
ΔG = −nF E (2.4.7)
cell

A spontaneous redox reaction is characterized by a negative value of ΔG°, which corresponds to a positive value of E°cell.

 Example 2.4.1

Suppose you want to prepare elemental bromine from bromide using the dichromate ion as an oxidant. Using the data in Table P2, calculate the free-energy change (ΔG°) for this redox reaction
under standard conditions. Is the reaction spontaneous?
Given: redox reaction
Asked for: ΔG for the reaction and spontaneity
o

Strategy:
A. From the relevant half-reactions and the corresponding values of E , write the overall reaction and calculate E .
o ∘
cell

B. Determine the number of electrons transferred in the overall reaction. Then use Equation 2.4.7 to calculate ΔG . If ΔG is negative, then the reaction is spontaneous.
o o

Solution

A
As always, the first step is to write the relevant half-reactions and use them to obtain the overall reaction and the magnitude of E . From Table P2, we can find the reduction and oxidation half-
o

reactions and corresponding E values:


o

2− + − 3+ ∘
cathode: C r2 O (aq) + 14 H (aq) + 6 e → 2C r (aq) + 7 H2 O(l) E = 1.36 V
7 cathode

− − ∘
anode: 2Br (aq) → Br2 (aq) + 2 e E = 1.09 V
anode

To obtain the overall balanced chemical equation, we must multiply both sides of the oxidation half-reaction by 3 to obtain the same number of electrons as in the reduction half-reaction,
remembering that the magnitude of E is not affected:
o

2.4.1 https://chem.libretexts.org/@go/page/48480
2− + − 3+ ∘
cathode: C r2 O (aq) + 14 H (aq) + 6 e → 2C r (aq) + 7 H2 O(l) E = 1.36 V
7 cathode

− − ∘
anode: 6Br (aq) → 3Br2 (aq) + 6 e E = 1.09 V
anode

2− − + 3+ ∘
overall: C r2 O (aq) + 6Br (aq) + 14 H (aq) → 2C r (aq) + 3Br2 (aq) + 7 H2 O(l) E = 0.27 V
7 cell

B
We can now calculate ΔG° using Equation 2.4.7. Because six electrons are transferred in the overall reaction, the value of n is 6:
∘ ∘
ΔG = −(n)(F )(E )
cell

= −(6 mole)[96, 485 J/(V ⋅ mol)(0.27 V)]

4
= −15.6  ×10  J
2−
= −156 kJ/mol Cr2 O
7

Thus ΔG is −168 kJ/mol for the reaction as written, and the reaction is spontaneous.
o

 Exercise 2.4.1
Use the data in Table P2 to calculate ΔG for the reduction of ferric ion by iodide:
o

3 + − 2 +
2 Fe (aq) + 2 I (aq) → 2 Fe (aq) + I (s)
2

Is the reaction spontaneous?

Answer
−44 kJ/mol I2; yes

I'm not a robot


reCAPTCHA
Privacy - Terms

About this page

Our systems have detected unusual traffic from your computer


network. This page checks to see if it's really you sending the
requests, and not a robot. Why did this happen?

IP address: 2604:a880:2:d0::2056:8001
Time: 2023-03-12T20:40:25Z
URL: https://www.youtube-nocookie.com/embed/49KhN73HVM8?
vq=hd1080

Relating G and Ecell: Relating G and Ecell(opens in new window) [youtu.be]

Potentials for the Sums of Half-Reactions


Although Table P2 list several half-reactions, many more are known. When the standard potential for a half-reaction is not available, we can use relationships between standard potentials and free
energy to obtain the potential of any other half-reaction that can be written as the sum of two or more half-reactions whose standard potentials are available. For example, the potential for the
reduction of Fe (aq) to Fe(s) is not listed in the table, but two related reductions are given:
3 +

3 + − 2 + ∘
Fe (aq) + e ⟶ Fe (aq) E = +0.77V (2.4.8)

2 + − ∘
Fe (aq) + 2 e ⟶ Fe(s) E = −0.45V (2.4.9)

Although the sum of these two half-reactions gives the desired half-reaction, we cannot simply add the potentials of two reductive half-reactions to obtain the potential of a third reductive half-
reaction because E is not a state function. However, because ΔG is a state function, the sum of the ΔG values for the individual reactions gives us ΔG for the overall reaction, which is
o o o o

proportional to both the potential and the number of electrons (n ) transferred. To obtain the value of E for the overall half-reaction, we first must add the values of ΔG (= −nF E ) for each
o o o

individual half-reaction to obtain ΔG for the overall half-reaction:


o

3 + − 2+ ∘
Fe (aq) + e → Fe (aq) ΔG = −(1)(F )(0.77 V)

2 + − ∘
Fe (aq) + 2 e → Fe(s) ΔG = −(2)(F )(−0.45 V)

3 + − ∘
Fe (aq) + 3 e → Fe(s) ΔG = [−(1)(F )(0.77 V)] + [−(2)(F )(−0.45 V)]

Solving the last expression for ΔG° for the overall half-reaction,

ΔG = F [(−0.77V ) + (−2)(−0.45V )] = F (0.13V ) (2.4.10)

Three electrons (n = 3 ) are transferred in the overall reaction, so substituting into Equation 2.4.7 and solving for E gives the following: o

∘ ∘
ΔG = −nF E
cell


F (0.13 V) = −(3)(F )(E )
cell

0.13 V

E =− = −0.043 V
3

2.4.2 https://chem.libretexts.org/@go/page/48480
This value of E is very different from the value that is obtained by simply adding the potentials for the two half-reactions (0.32 V) and even has the opposite sign.
o

Values of E for half-reactions cannot be added to give E for the sum of the half-reactions; only values of ΔG
o o o
= −nF E

cell
for half-reactions
can be added.

The Relationship between Cell Potential & the Equilibrium Constant


We can use the relationship between ΔG and the equilibrium constant K , to obtain a relationship between

E

cell
and K . Recall that for a general reaction of the type aA + bB → cC + dD , the
standard free-energy change and the equilibrium constant are related by the following equation:

ΔG° = −RT ln K (2.4.11)

Given the relationship between the standard free-energy change and the standard cell potential (Equation 2.4.7), we can write

−nF E = −RT ln K (2.4.12)
cell

Rearranging this equation,


RT

E =( ) ln K (2.4.13)
cell
nF

For T = 298 K , Equation 2.4.13 can be simplified as follows:


RT

E =( ) ln K (2.4.14)
cell
nF

[8.314 J/(mol ⋅ K)(298 K)]


=[ ] 2.303 log K (2.4.15)
n[96, 485 J/(V ⋅ mol)]

0.0592 V
=( ) log K (2.4.16)
n

Thus E ∘
cell
is directly proportional to the logarithm of the equilibrium constant. This means that large equilibrium constants correspond to large positive values of E ∘
cell
and vice versa.

 Example 2.4.2

Use the data in Table P2 to calculate the equilibrium constant for the reaction of metallic lead with PbO2 in the presence of sulfate ions to give PbSO4 under standard conditions. (This reaction
occurs when a car battery is discharged.) Report your answer to two significant figures.
Given: redox reaction
Asked for: K

Strategy:
A. Write the relevant half-reactions and potentials. From these, obtain the overall reaction and E . o
cell

B. Determine the number of electrons transferred in the overall reaction. Use Equation 2.4.16 to solve for log K and then K .

Solution
A The relevant half-reactions and potentials from Table P2 are as follows:
2− + − ∘
cathode: PbO2 (s) + SO (aq) + 4 H (aq) + 2 e → PbSO4 (s) + 2 H2 O(l) E = 1.69 V
4 cathode

2− − ∘
anode: Pb(s) + SO (aq) → PbSO4 (s) + 2 e E = −0.36 V
4 anode

2− + ∘
overall: Pb(s) + PbO2 (s) + 2SO (aq) + 4 H (aq) → 2PbSO4 (s) + 2 H2 O(l) E = 2.05 V
4 cell

B Two electrons are transferred in the overall reaction, so n = 2 . Solving Equation 2.4.16 for log K and inserting the values of n and E , o


nE 2(2.05 V)
log K = = = 69.37
0.0591 V 0.0591 V

69
K = 2.3 × 10

Thus the equilibrium lies far to the right, favoring a discharged battery (as anyone who has ever tried unsuccessfully to start a car after letting it sit for a long time will know).

 Exercise 2.4.2

Use the data in Table P2 to calculate the equilibrium constant for the reaction of Sn
2 +
(aq) with oxygen to produce Sn
4 +
(aq) and water under standard conditions. Report your answer to two
significant figures. The reaction is as follows:
2 + + 4 +
2 Sn (aq) + O (g) + 4 H (aq) −
↽⇀
− 2 Sn (aq) + 2 H O(l)
2 2

Answer
72
5.7 × 10

Figure 2.4.1 summarizes the relationships that we have developed based on properties of the system—that is, based on the equilibrium constant, standard free-energy change, and standard cell
potential—and the criteria for spontaneity (ΔG° < 0). Unfortunately, these criteria apply only to systems in which all reactants and products are present in their standard states, a situation that is
seldom encountered in the real world. A more generally useful relationship between cell potential and reactant and product concentrations, as we are about to see, uses the relationship between ΔG
and the reaction quotient Q.

2.4.3 https://chem.libretexts.org/@go/page/48480
Figure 2.4.1 : The Relationships among Criteria for Thermodynamic Spontaneity. The three properties of a system that can be used to predict the spontaneity of a redox reaction under standard
conditions are K, ΔG°, and E°cell. If we know the value of one of these quantities, then these relationships enable us to calculate the value of the other two. The signs of ΔG° and E°cell and the
magnitude of K determine the direction of spontaneous reaction under standard conditions. (CC BY-NC-SA; Anonymous by request)
If delta G is less than zero, E is greater than zero and K is greater than 1 then the direction of the reaction is spontaneous in forward direction. If delta G is greater than zero, E is less than zero and K is
less than one then the direction of reaction is spontaneous in reverse direction. If delta G is zero, E is zero and k is one that there is no net reaction and the system is at equilibrium .

I'm not a robot


reCAPTCHA
Privacy - Terms

About this page

Our systems have detected unusual traffic from your computer


network. This page checks to see if it's really you sending the
requests, and not a robot. Why did this happen?

IP address: 2604:a880:2:d0::2056:8001
Time: 2023-03-12T20:40:25Z
URL: https://www.youtube-nocookie.com/embed/zeeAXleT1c0?
vq=hd1080

Electrode Potentials and ECell: Electrode Potentials and Ecell(opens in new window) [youtu.be]

Summary
A coulomb (C) relates electrical potential, expressed in volts, and energy, expressed in joules. The current generated from a redox reaction is measured in amperes (A), where 1 A is defined as the
flow of 1 C/s past a given point. The faraday (F) is Avogadro’s number multiplied by the charge on an electron and corresponds to the charge on 1 mol of electrons. The product of the cell potential
and the total charge is the maximum amount of energy available to do work, which is related to the change in free energy that occurs during the chemical process. Adding together the ΔG values for
the half-reactions gives ΔG for the overall reaction, which is proportional to both the potential and the number of electrons (n) transferred. Spontaneous redox reactions have a negative ΔG and
therefore a positive Ecell. Because the equilibrium constant K is related to ΔG, E°cell and K are also related. Large equilibrium constants correspond to large positive values of E°.

2.4: E, ΔG, and K (The Big Triangle of Chemistry) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
20.5: Gibbs Energy and Redox Reactions is licensed CC BY-NC-SA 3.0.

2.4.4 https://chem.libretexts.org/@go/page/48480
2.4.5 https://chem.libretexts.org/@go/page/48480
2.5: Cell Potential as a Function of Concentrations
 Learning Objectives
Relate cell potentials to Gibbs energy changes
Use the Nernst equation to determine cell potentials at nonstandard conditions
Perform calculations that involve converting between cell potentials, free energy changes, and equilibrium constants

The Nernst Equation enables the determination of cell potential under non-standard conditions. It relates the measured cell
potential to the reaction quotient and allows the accurate determination of equilibrium constants (including solubility constants).

The Effect of Concentration on Cell Potential: The Nernst Equation


Recall that the actual free-energy change for a reaction under nonstandard conditions, ΔG, is given as follows:
\[\Delta{G} = \Delta{G°} + RT \ln Q \label{Eq1} \]
We also know that ΔG = −nF E cell(under non-standard confitions) and ΔG
o
= −nF E
cell
o
(under standard conditions).
Substituting these expressions into Equation ??? , we obtain
o
−nF Ecell = −nF E + RT ln Q (2.5.1)
cell

Dividing both sides of this equation by −nF ,


RT

Ecell = E −( ) ln Q (2.5.2)
cell
nF

Equation 2.5.2 is called the Nernst equation, after the German physicist and chemist Walter Nernst (1864–1941), who first
derived it. The Nernst equation is arguably the most important relationship in electrochemistry. When a redox reaction is at
equilibrium (ΔG = 0 ), then Equation 2.5.2 reduces to Equation 2.5.3 and 2.5.4 because Q = K , and there is no net transfer of
electrons (i.e., Ecell = 0).
RT

Ecell = E −( ) ln K = 0 (2.5.3)
cell
nF

since
RT

E =( ) ln K (2.5.4)
cell
nF

Substituting the values of the constants into Equation 2.5.2 with T = 298 K and converting to base-10 logarithms give the
relationship of the actual cell potential (Ecell), the standard cell potential (E°cell), and the reactant and product concentrations at
room temperature (contained in Q):
0.0591 V

Ecell = E −( ) log Q (2.5.5)
cell
n

 The Power of the Nernst Equation


The Nernst Equation (2.5.2) can be used to determine the value of Ecell, and thus the direction of spontaneous reaction, for any
redox reaction under any conditions.

Equation 2.5.5 allows us to calculate the potential associated with any electrochemical cell at 298 K for any combination of
reactant and product concentrations under any conditions. We can therefore determine the spontaneous direction of any redox
reaction under any conditions, as long as we have tabulated values for the relevant standard electrode potentials. Notice in Equation
2.5.5 that the cell potential changes by 0.0591/n V for each 10-fold change in the value of Q because log 10 = 1.

2.5.1 https://chem.libretexts.org/@go/page/48481
 Example 2.5.1

The following reaction proceeds spontaneously under standard conditions because E°cell > 0 (which means that ΔG° < 0):
4 + – 3 + ∘
2 Ce (aq) + 2 Cl (aq) ⟶ 2 Ce (aq) + Cl (g)  E = 0.25 V
2 cell

Calculate E for this reaction under the following nonstandard conditions and determine whether it will occur spontaneously:
cell

[Ce4+] = 0.013 M, [Ce3+] = 0.60 M, [Cl−] = 0.0030 M, P = 1.0 atm, and T = 25°C.
Cl2

Given: balanced redox reaction, standard cell potential, and nonstandard conditions
Asked for: cell potential

Strategy:
Determine the number of electrons transferred during the redox process. Then use the Nernst equation to find the cell potential
under the nonstandard conditions.

Solution
We can use the information given and the Nernst equation to calculate Ecell. Moreover, because the temperature is 25°C (298
K), we can use Equation 2.5.5 instead of Equation 2.5.2. The overall reaction involves the net transfer of two electrons:
4+ − 3+
2C e + 2e → 2C e
(aq) (aq)

− −
2C l → C l2(g) + 2 e
(aq)

so n = 2. Substituting the concentrations given in the problem, the partial pressure of Cl2, and the value of E°cell into Equation
2.5.5,

0.0591 V

Ecell = E −( ) log Q
cell
n

3+ 2
0.0591 V [Ce ] PCl2
= 0.25 V − ( ) log( )
4+ 2 − 2
2 [C e ] [C l ]

= 0.25 V − [(0.0296 V)(8.37)] = 0.00 V

Thus the reaction will not occur spontaneously under these conditions (because E = 0 V and ΔG = 0). The composition
specified is that of an equilibrium mixture

 Exercise 2.5.1

Molecular oxygen will not oxidize M nO to permanganate via the reaction


2

− −
4 MnO (s) + 3 O (g) + 4 OH (aq) ⟶ 4 MnO 4 (aq) + 2 H O(l) E °cell = −0.20 V
2 2 2

Calculate E cellfor the reaction under the following nonstandard conditions and decide whether the reaction will occur
spontaneously: pH 10, P = 0.20 atm, [MNO4−] = 1.0 × 10−4 M, and T = 25°C.
O2

Answer
Ecell = −0.22 V; the reaction will not occur spontaneously.

Applying the Nernst equation to a simple electrochemical cell such as the Zn/Cu cell allows us to see how the cell voltage varies as
the reaction progresses and the concentrations of the dissolved ions change. Recall that the overall reaction for this cell is as
follows:
2+ 2+
Zn(s) + C u (aq) → Z n (aq) + C u(s) E°cell = 1.10V (2.5.6)

2.5.2 https://chem.libretexts.org/@go/page/48481
The reaction quotient is therefore Q = [Z n ]/[C u ]. Suppose that the cell initially contains 1.0 M Cu2+ and 1.0 × 10−6 M Zn2+.
2+ 2+

The initial voltage measured when the cell is connected can then be calculated from Equation 2.5.5:
2+
0.0591 V [Zn ]

Ecell = E −( ) log (2.5.7)
cell 2+
n [C u ]

−6
0.0591 V 1.0 × 10
= 1.10 V − ( ) log( ) = 1.28 V (2.5.8)
2 1.0

Thus the initial voltage is greater than E° because Q < 1 . As the reaction proceeds, [Zn2+] in the anode compartment increases as
the zinc electrode dissolves, while [Cu2+] in the cathode compartment decreases as metallic copper is deposited on the electrode.
During this process, the ratio Q = [Zn2+]/[Cu2+] steadily increases, and the cell voltage therefore steadily decreases. Eventually,
[Zn2+] = [Cu2+], so Q = 1 and Ecell = E°cell. Beyond this point, [Zn2+] will continue to increase in the anode compartment, and
[Cu2+] will continue to decrease in the cathode compartment. Thus the value of Q will increase further, leading to a further decrease
in Ecell. When the concentrations in the two compartments are the opposite of the initial concentrations (i.e., 1.0 M Zn2+ and 1.0 ×
10−6 M Cu2+), Q = 1.0 × 106, and the cell potential will be reduced to 0.92 V.

Figure 2.5.1 : The Variation of Ecell with Log Q for a Zn/Cu Cell. Initially, log Q < 0, and the voltage of the cell is greater than
E°cell. As the reaction progresses, log Q increases, and Ecell decreases. When [Zn2+] = [Cu2+], log Q = 0 and Ecell = E°cell = 1.10 V.
As long as the electrical circuit remains intact, the reaction will continue, and log Q will increase until Q = K and the cell voltage
reaches zero. At this point, the system will have reached equilibrium.
The variation of Ecell with log Q over this range is linear with a slope of −0.0591/n, as illustrated in Figure 2.5.1. As the reaction
proceeds still further, Q continues to increase, and Ecell continues to decrease. If neither of the electrodes dissolves completely,
thereby breaking the electrical circuit, the cell voltage will eventually reach zero. This is the situation that occurs when a battery is
“dead.” The value of Q when Ecell = 0 is calculated as follows:
0.0591 V

Ecell = E −( ) log Q = 0 (2.5.9)
cell
n

0.0591 V

E =( ) log Q (2.5.10)
n

E n (1.10 V)(2)
log Q = = = 37.23 (2.5.11)
0.0591 V 0.0591 V
37.23 37
Q = 10 = 1.7 × 10 (2.5.12)

Recall that at equilibrium, Q = K . Thus the equilibrium constant for the reaction of Zn metal with Cu2+ to give Cu metal and Zn2+
is 1.7 × 1037 at 25°C.

2.5.3 https://chem.libretexts.org/@go/page/48481
I'm not a robot
reCAPTCHA
Privacy - Terms

About this page

Our systems have detected unusual traffic from your


computer network This page checks to see if it's really you
The Nernst Equation: The Nernst Equation (opens in new window) [youtu.be]

Concentration Cells
A voltage can also be generated by constructing an electrochemical cell in which each compartment contains the same redox active
solution but at different concentrations. The voltage is produced as the concentrations equilibrate. Suppose, for example, we have a
cell with 0.010 M AgNO3 in one compartment and 1.0 M AgNO3 in the other. The cell diagram and corresponding half-reactions
are as follows:
+ +
Ag(s) | Ag (aq, 0.010 M ) || Ag (aq, 1.0 M ) | Ag(s) (2.5.13)

cathode:
+ −
Ag (aq, 1.0 M ) + e → Ag(s) (2.5.14)

anode:
+ −
Ag(s) → Ag (aq, 0.010 M ) + e (2.5.15)

Overall
+ +
Ag (aq, 1.0 M ) → Ag (aq, 0.010 M ) (2.5.16)

As the reaction progresses, the concentration of Ag will increase in the left (oxidation) compartment as the silver electrode
+

dissolves, while the Ag concentration in the right (reduction) compartment decreases as the electrode in that compartment gains
+

mass. The total mass of Ag(s) in the cell will remain constant, however. We can calculate the potential of the cell using the Nernst
equation, inserting 0 for E°cell because E°cathode = −E°anode:
0.0591 V

Ecell = E −( ) log Q
cell
n

0.0591 V 0.010
= 0 −( ) log( )
1 1.0

= 0.12 V

An electrochemical cell of this type, in which the anode and cathode compartments are identical except for the concentration of a
reactant, is called a concentration cell. As the reaction proceeds, the difference between the concentrations of Ag+ in the two
compartments will decrease, as will Ecell. Finally, when the concentration of Ag+ is the same in both compartments, equilibrium
will have been reached, and the measured potential difference between the two compartments will be zero (Ecell = 0).

 Example 2.5.2
Calculate the voltage in a galvanic cell that contains a manganese electrode immersed in a 2.0 M solution of MnCl2 as the
cathode, and a manganese electrode immersed in a 5.2 × 10−2 M solution of MnSO4 as the anode (T = 25°C).
Given: galvanic cell, identities of the electrodes, and solution concentrations

2.5.4 https://chem.libretexts.org/@go/page/48481
Asked for: voltage

Strategy:
A. Write the overall reaction that occurs in the cell.
B. Determine the number of electrons transferred. Substitute this value into the Nernst equation to calculate the voltage.

Solution
A This is a concentration cell, in which the electrode compartments contain the same redox active substance but at different
concentrations. The anions (Cl− and SO42−) do not participate in the reaction, so their identity is not important. The overall
reaction is as follows:
2 + 2 + −2
Mn (aq, 2.0 M ) → Mn (aq, 5.2 × 10 M)

B For the reduction of Mn2+(aq) to Mn(s), n = 2. We substitute this value and the given Mn2+ concentrations into Equation
2.5.5:

0.0591 V

Ecell = E −( ) log Q
cell
n

−2
0.0591 V 5.2 × 10
= 0 V−( ) log( )
2 2.0

= 0.047 V

Thus manganese will dissolve from the electrode in the compartment that contains the more dilute solution and will be
deposited on the electrode in the compartment that contains the more concentrated solution.

 Exercise 2.5.2

Suppose we construct a galvanic cell by placing two identical platinum electrodes in two beakers that are connected by a salt
bridge. One beaker contains 1.0 M HCl, and the other a 0.010 M solution of Na2SO4 at pH 7.00. Both cells are in contact with
the atmosphere, with P = 0.20 atm. If the relevant electrochemical reaction in both compartments is the four-electron
O2

reduction of oxygen to water:


+ −
O (g) + 4 H (aq) + 4 e → 2 H O(l)
2 2

What will be the potential when the circuit is closed?

Answer
0.41 V

Using Cell Potentials to Measure Solubility Products


Because voltages are relatively easy to measure accurately using a voltmeter, electrochemical methods provide a convenient way to
determine the concentrations of very dilute solutions and the solubility products (K ) of sparingly soluble substances. As you
sp

learned previously, solubility products can be very small, with values of less than or equal to 10−30. Equilibrium constants of this
magnitude are virtually impossible to measure accurately by direct methods, so we must use alternative methods that are more
sensitive, such as electrochemical methods.

2.5.5 https://chem.libretexts.org/@go/page/48481
Figure 2.5.1 : A Galvanic ("Concentration") Cell for Measuring the Solubility Product of AgCl. One compartment contains a silver
wire inserted into a 1.0 M solution of Ag+, and the other compartment contains a silver wire inserted into a 1.0 M Cl− solution
saturated with AgCl. The potential due to the difference in [Ag+] between the two cells can be used to determine K . (CC BY-NC-sp

SA; Anonymous by request)


To understand how an electrochemical cell is used to measure a solubility product, consider the cell shown in Figure 2.5.1, which
is designed to measure the solubility product of silver chloride:
+ −
Ksp = [ Ag ][ Cl ].

In one compartment, the cell contains a silver wire inserted into a 1.0 M solution of Ag+; the other compartment contains a silver
wire inserted into a 1.0 M Cl− solution saturated with AgCl. In this system, the Ag+ ion concentration in the first compartment
equals Ksp. We can see this by dividing both sides of the equation for Ksp by [Cl−] and substituting:
Ksp
+
[ Ag ] =

[ Cl ]

Ksp
= = Ksp .
1.0

The overall cell reaction is as follows:


Ag+(aq, concentrated) → Ag+(aq, dilute)
Thus the voltage of the concentration cell due to the difference in [Ag+] between the two cells is as follows:
+
0.0591 V [Ag ]dilute
Ecell = 0 V − ( ) log( )
+
1 [Ag ]concentrated

Ksp
= −0.0591 V  log( )
1.0

= −0.0591 V  log Ksp (2.5.17)

By closing the circuit, we can measure the potential caused by the difference in [Ag+] in the two cells. In this case, the
experimentally measured voltage of the concentration cell at 25°C is 0.580 V. Solving Equation 2.5.17 for K , sp

−Ecell −0.580 V
log Ksp = = = −9.81
0.0591 V 0.0591 V

−10
Ksp = 1.5 × 10

Thus a single potential measurement can provide the information we need to determine the value of the solubility product of a
sparingly soluble salt.

2.5.6 https://chem.libretexts.org/@go/page/48481
 Example 2.5.3: Solubility of lead(II) sulfate
To measure the solubility product of lead(II) sulfate (PbSO4) at 25°C, you construct a galvanic cell like the one shown in
Figure 2.5.1, which contains a 1.0 M solution of a very soluble Pb2+ salt [lead(II) acetate trihydrate] in one compartment that is
connected by a salt bridge to a 1.0 M solution of Na2SO4 saturated with PbSO4 in the other. You then insert a Pb electrode into
each compartment and close the circuit. Your voltmeter shows a voltage of 230 mV. What is Ksp for PbSO4? Report your
answer to two significant figures.
Given: galvanic cell, solution concentrations, electrodes, and voltage
Asked for: Ksp

Strategy:
A. From the information given, write the equation for Ksp. Express this equation in terms of the concentration of Pb2+.
B. Determine the number of electrons transferred in the electrochemical reaction. Substitute the appropriate values into
Equation ??? 2 and solve for Ksp.

Solution
A You have constructed a concentration cell, with one compartment containing a 1.0 M solution of Pb and the other
2 +

containing a dilute solution of Pb2+ in 1.0 M Na2SO4. As for any concentration cell, the voltage between the two compartments
can be calculated using the Nernst equation. The first step is to relate the concentration of Pb2+ in the dilute solution to Ksp:
2+ 2−
[Pb ][SO ] = Ksp
4

Ksp Ksp
2+
[Pb ] = = = Ksp
2−
[SO ] 1.0 M
4

B The reduction of Pb2+ to Pb is a two-electron process and proceeds according to the following reaction:
Pb2+(aq, concentrated) → Pb2+(aq, dilute)
so
0.0591

Ecell = E −( ) log Q
cell
n
2+
0.0591 V [Pb ]dilute Ksp
0.230 V = 0 V − ( ) log( ) = −0.0296 V log( )
2+
2 [Pb ]concentrated 1.0

−7.77 = log Ksp

−8
1.7 × 10 = Ksp

 Exercise 2.5.3

A concentration cell similar to the one described in Example 2.5.3 contains a 1.0 M solution of lanthanum nitrate [La(NO3)3]
in one compartment and a 1.0 M solution of sodium fluoride saturated with LaF3 in the other. A metallic La strip is inserted
into each compartment, and the circuit is closed. The measured potential is 0.32 V. What is the Ksp for LaF3? Report your
answer to two significant figures.

Answer
5.7 × 10−17

Using Cell Potentials to Measure Concentrations


Another use for the Nernst equation is to calculate the concentration of a species given a measured potential and the concentrations
of all the other species. We saw an example of this in Example 2.5.3, in which the experimental conditions were defined in such a

2.5.7 https://chem.libretexts.org/@go/page/48481
way that the concentration of the metal ion was equal to Ksp. Potential measurements can be used to obtain the concentrations of
dissolved species under other conditions as well, which explains the widespread use of electrochemical cells in many analytical
devices. Perhaps the most common application is in the determination of [H+] using a pH meter, as illustrated below.

 Example 2.5.4: Measuring pH

Suppose a galvanic cell is constructed with a standard Zn/Zn2+ couple in one compartment and a modified hydrogen electrode
in the second compartment. The pressure of hydrogen gas is 1.0 atm, but [H+] in the second compartment is unknown. The cell
diagram is as follows:
2 + +
Zn(s)| Zn (aq, 1.0 M )|| H (aq, ? M )| H (g, 1.0 atm)|P t(s)
2

What is the pH of the solution in the second compartment if the measured potential in the cell is 0.26 V at 25°C?
Given: galvanic cell, cell diagram, and cell potential
Asked for: pH of the solution

Strategy:
A. Write the overall cell reaction.
B. Substitute appropriate values into the Nernst equation and solve for −log[H+] to obtain the pH.

Solution
A Under standard conditions, the overall reaction that occurs is the reduction of protons by zinc to give H2 (note that Zn lies
below H2 in Table P2):
Zn(s) + 2H2+(aq) → Zn2+(aq) + H2(g) E°=0.76 V
B By substituting the given values into the simplified Nernst equation (Equation 2.5.5 ), we can calculate [H+] under
nonstandard conditions:
2+
0.0591 V [Zn ] PH
2

Ecell = E −( ) log( )
cell + 2
n [H ]

0.0591 V (1.0)(1.0)
0.26 V = 0.76 V − ( ) log( )
+ 2
2 [H ]

1
+ −2 +
16.9 = log( ) = log[ H ] = (−2) log[ H ]
+ 2
[H ]
+
8.46 = − log[ H ]

8.5 = pH

Thus the potential of a galvanic cell can be used to measure the pH of a solution.

 Exercise 2.5.4

Suppose you work for an environmental laboratory and you want to use an electrochemical method to measure the
concentration of Pb2+ in groundwater. You construct a galvanic cell using a standard oxygen electrode in one compartment
(E°cathode = 1.23 V). The other compartment contains a strip of lead in a sample of groundwater to which you have added
sufficient acetic acid, a weak organic acid, to ensure electrical conductivity. The cell diagram is as follows:
2+ +
P b(s) ∣ P b (aq, ?M ) ∥ H (aq), 1.0M ∣ O2 (g, 1.0atm) ∣ P t(s)

When the circuit is closed, the cell has a measured potential of 1.62 V. Use Table P2 to determine the concentration of Pb2+ in
the groundwater.

Answer

2.5.8 https://chem.libretexts.org/@go/page/48481
−9
1.2 × 10 M

Summary
The Nernst equation can be used to determine the direction of spontaneous reaction for any redox reaction in aqueous solution. The
Nernst equation allows us to determine the spontaneous direction of any redox reaction under any reaction conditions from values
of the relevant standard electrode potentials. Concentration cells consist of anode and cathode compartments that are identical
except for the concentrations of the reactant. Because ΔG = 0 at equilibrium, the measured potential of a concentration cell is zero
at equilibrium (the concentrations are equal). A galvanic cell can also be used to measure the solubility product of a sparingly
soluble substance and calculate the concentration of a species given a measured potential and the concentrations of all the other
species.

2.5: Cell Potential as a Function of Concentrations is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.
20.6: Cell Potential Under Nonstandard Conditions is licensed CC BY-NC-SA 3.0.

2.5.9 https://chem.libretexts.org/@go/page/48481
2.6: Batteries- Producing Electricity Through Chemical Reactions
Because galvanic cells can be self-contained and portable, they can be used as batteries and fuel cells. A battery (storage cell) is a galvanic cell (or
a series of galvanic cells) that contains all the reactants needed to produce electricity. In contrast, a fuel cell is a galvanic cell that requires a constant
external supply of one or more reactants to generate electricity. In this section, we describe the chemistry behind some of the more common types of
batteries and fuel cells.

Batteries
There are two basic kinds of batteries: disposable, or primary, batteries, in which the electrode reactions are effectively irreversible and which cannot
be recharged; and rechargeable, or secondary, batteries, which form an insoluble product that adheres to the electrodes. These batteries can be
recharged by applying an electrical potential in the reverse direction. The recharging process temporarily converts a rechargeable battery from a
galvanic cell to an electrolytic cell.
Batteries are cleverly engineered devices that are based on the same fundamental laws as galvanic cells. The major difference between batteries and
the galvanic cells we have previously described is that commercial batteries use solids or pastes rather than solutions as reactants to maximize the
electrical output per unit mass. The use of highly concentrated or solid reactants has another beneficial effect: the concentrations of the reactants and
the products do not change greatly as the battery is discharged; consequently, the output voltage remains remarkably constant during the discharge
process. This behavior is in contrast to that of the Zn/Cu cell, whose output decreases logarithmically as the reaction proceeds (Figure 2.6.1). When
a battery consists of more than one galvanic cell, the cells are usually connected in series—that is, with the positive (+) terminal of one cell
connected to the negative (−) terminal of the next, and so forth. The overall voltage of the battery is therefore the sum of the voltages of the
individual cells.

Figure 2.6.1 : Three Kinds of Primary (Nonrechargeable) Batteries. (a) A Leclanché dry cell is actually a “wet cell,” in which the electrolyte is an
acidic water-based paste containing MnO2, NH4Cl, ZnCl2, graphite, and starch. Though inexpensive to manufacture, the cell is not very efficient in
producing electrical energy and has a limited shelf life. (b) In a button battery, the anode is a zinc–mercury amalgam, and the cathode can be either
HgO (shown here) or Ag2O as the oxidant. Button batteries are reliable and have a high output-to-mass ratio, which allows them to be used in
applications such as calculators and watches, where their small size is crucial. (c) A lithium–iodine battery consists of two cells separated by a
metallic nickel mesh that collects charge from the anodes. The anode is lithium metal, and the cathode is a solid complex of I2. The electrolyte is a
layer of solid LiI that allows Li+ ions to diffuse from the cathode to the anode. Although this type of battery produces only a relatively small current,
it is highly reliable and long-lived.

The major difference between batteries and the galvanic cells is that commercial typically batteries use solids or pastes rather than solutions as
reactants to maximize the electrical output per unit mass. An obvious exception is the standard car battery which used solution phase chemistry.

Leclanché Dry Cell


The dry cell, by far the most common type of battery, is used in flashlights, electronic devices such as the Walkman and Game Boy, and many other
devices. Although the dry cell was patented in 1866 by the French chemist Georges Leclanché and more than 5 billion such cells are sold every year,
the details of its electrode chemistry are still not completely understood. In spite of its name, the Leclanché dry cell is actually a “wet cell”: the
electrolyte is an acidic water-based paste containing M nO , N H C l, ZnC l , graphite, and starch (part (a) in Figure 2.6.1). The half-reactions at
2 4 2

the anode and the cathode can be summarized as follows:


cathode (reduction):
+ −
2 MnO (s) + 2 NH (aq) + 2 e ⟶ Mn O (s) + 2 NH (aq) + H O(l)
2 4 2 3 3 2

anode (oxidation):
2 + −
Zn(s) ⟶ Zn (aq) + 2 e

2.6.1 https://chem.libretexts.org/@go/page/48482
The Zn ions formed by the oxidation of Zn(s) at the anode react with NH formed at the cathode and Cl ions present in solution, so the overall
2 +

3

cell reaction is as follows:


overall reaction:
2 MnO (s) + 2 NH Cl(aq) + Zn(s) ⟶ Mn O (s) + Zn(NH ) Cl (s) + H O(l) (2.6.1)
2 4 2 3 3 2 2 2

The dry cell produces about 1.55 V and is inexpensive to manufacture. It is not, however, very efficient in producing electrical energy because only
the relatively small fraction of the MnO that is near the cathode is actually reduced and only a small fraction of the zinc cathode is actually
2

consumed as the cell discharges. In addition, dry cells have a limited shelf life because the Zn anode reacts spontaneously with NH Cl in the 4

electrolyte, causing the case to corrode and allowing the contents to leak out.
Close up of a hand holding one double <span class=
AA
battery" style="width: 287px; height: 215px;" width="287px" height="215px" data-cke-saved-src="/@api/deki/files/16647/battery.jpg"
src="/@api/deki/files/16647/battery.jpg" data-quail-id="34">
Source: Photo courtesy of Mitchclanky2008, www.flickr.com/photos/25597837@N05/2422765479/.
The alkaline battery is essentially a Leclanché cell adapted to operate under alkaline, or basic, conditions. The half-reactions that occur in an
alkaline battery are as follows:
cathode (reduction)
− −
2 MnO (s) + H O(l) + 2 e ⟶ Mn O (s) + 2 OH (aq)
2 2 2 3

anode (oxidation):
− −
Zn(s) + 2 OH (aq) ⟶ ZnO(s) + H O(l) + 2 e
2

overall reaction:

Zn(s) + 2 MnO (s) ⟶ ZnO(s) + Mn O (s)


2 2 3

This battery also produces about 1.5 V, but it has a longer shelf life and more constant output voltage as the cell is discharged than the Leclanché dry
cell. Although the alkaline battery is more expensive to produce than the Leclanché dry cell, the improved performance makes this battery more
cost-effective.

Button Batteries
Although some of the small button batteries used to power watches, calculators, and cameras are miniature alkaline cells, most are based on a
completely different chemistry. In these "button" batteries, the anode is a zinc–mercury amalgam rather than pure zinc, and the cathode uses either
HgO or Ag O as the oxidant rather than MnO in Figure 2.6.1b).
2 2

Button batteries. (Gerhard H Wrodnigg via Wikipedia)


The cathode, anode and overall reactions and cell output for these two types of button batteries are as follows (two half-reactions occur at the anode,
but the overall oxidation half-reaction is shown):
cathode (mercury battery):
− −
HgO(s) + H O(l) + 2 e ⟶ Hg(l) + 2 OH (aq)
2

Anode (mercury battery):


− −
Zn + 2 OH ⟶ ZnO + H O + 2 e
2

overall reaction (mercury battery):

Zn(s) + 2 HgO(s) ⟶ 2 Hg(l) + ZnO(s)

with E = 1.35 V .
cell

cathode reaction (silver battery):


− −
Ag O(s) + H O(l) + 2 e ⟶ 2 Ag(s) + 2 OH (aq)
2 2

2.6.2 https://chem.libretexts.org/@go/page/48482
anode (silver battery):
− −
Zn + 2 OH ⟶ ZnO + H O + 2 e
2

Overall reaction (silver battery):

Zn(s) + 2 Ag O(s) ⟶ 2 Ag(s) + ZnO(s)


2

with E cell = 1.6 V .


The major advantages of the mercury and silver cells are their reliability and their high output-to-mass ratio. These factors make them ideal for
applications where small size is crucial, as in cameras and hearing aids. The disadvantages are the expense and the environmental problems caused
by the disposal of heavy metals, such as Hg and Ag.

Lithium–Iodine Battery
None of the batteries described above is actually “dry.” They all contain small amounts of liquid water, which adds significant mass and causes
potential corrosion problems. Consequently, substantial effort has been expended to develop water-free batteries. One of the few commercially
successful water-free batteries is the lithium–iodine battery. The anode is lithium metal, and the cathode is a solid complex of I . Separating them
2

is a layer of solid LiI, which acts as the electrolyte by allowing the diffusion of Li+ ions. The electrode reactions are as follows:
cathode (reduction):
− −
I2(s) + 2 e → 2I (LiI)
(2.6.2)

anode (oxidation):
+ −
2Li(s) → 2Li + 2e (2.6.3)
(LiI)

overall:

2Li(s) + I2(s) → 2Li I(s) (2.6.4)

with E cell = 3.5 V

Cardiac pacemaker: An x-ray of a patient showing the location and size of a pacemaker powered by a lithium–iodine battery.
As shown in part (c) in Figure 2.6.1, a typical lithium–iodine battery consists of two cells separated by a nickel metal mesh that collects charge from
the anode. Because of the high internal resistance caused by the solid electrolyte, only a low current can be drawn. Nonetheless, such batteries have
proven to be long-lived (up to 10 yr) and reliable. They are therefore used in applications where frequent replacement is difficult or undesirable, such
as in cardiac pacemakers and other medical implants and in computers for memory protection. These batteries are also used in security transmitters
and smoke alarms. Other batteries based on lithium anodes and solid electrolytes are under development, using T iS , for example, for the cathode.
2

Dry cells, button batteries, and lithium–iodine batteries are disposable and cannot be recharged once they are discharged. Rechargeable batteries, in
contrast, offer significant economic and environmental advantages because they can be recharged and discharged numerous times. As a result,
manufacturing and disposal costs drop dramatically for a given number of hours of battery usage. Two common rechargeable batteries are the
nickel–cadmium battery and the lead–acid battery, which we describe next.

Nickel–Cadmium (NiCad) Battery


The nickel–cadmium, or NiCad, battery is used in small electrical appliances and devices like drills, portable vacuum cleaners, and AM/FM digital
tuners. It is a water-based cell with a cadmium anode and a highly oxidized nickel cathode that is usually described as the nickel(III) oxo-hydroxide,
NiO(OH). As shown in Figure 2.6.2, the design maximizes the surface area of the electrodes and minimizes the distance between them, which
decreases internal resistance and makes a rather high discharge current possible.

2.6.3 https://chem.libretexts.org/@go/page/48482
Figure 2.6.2 : The Nickel–Cadmium (NiCad) Battery, a Rechargeable Battery. NiCad batteries contain a cadmium anode and a highly oxidized nickel
cathode. This design maximizes the surface area of the electrodes and minimizes the distance between them, which gives the battery both a high
discharge current and a high capacity.
The electrode reactions during the discharge of a N iC ad battery are as follows:
cathode (reduction):
− −
2N iO(OH )(s) + 2 H2 O(l) + 2 e → 2N i(OH )2(s) + 2OH (2.6.5)
(aq)

anode (oxidation):
− −
C d(s) + 2OH → C d(OH )2(s) + 2 e (2.6.6)
(aq)

overall:
C d(s) + 2N iO(OH )(s) + 2 H2 O(l) → C d(OH )2(s) + 2N i(OH )2(s) (2.6.7)

Ecell = 1.4V

Because the products of the discharge half-reactions are solids that adhere to the electrodes [Cd(OH)2 and 2Ni(OH)2], the overall reaction is readily
reversed when the cell is recharged. Although NiCad cells are lightweight, rechargeable, and high capacity, they have certain disadvantages. For
example, they tend to lose capacity quickly if not allowed to discharge fully before recharging, they do not store well for long periods when fully
charged, and they present significant environmental and disposal problems because of the toxicity of cadmium.
A variation on the NiCad battery is the nickel–metal hydride battery (NiMH) used in hybrid automobiles, wireless communication devices, and
mobile computing. The overall chemical equation for this type of battery is as follows:
\[NiO(OH)_{(s)} + MH \rightarrow Ni(OH)_{2(s)} + M_{(s)} \label{Eq16} \]
The NiMH battery has a 30%–40% improvement in capacity over the NiCad battery; it is more environmentally friendly so storage, transportation,
and disposal are not subject to environmental control; and it is not as sensitive to recharging memory. It is, however, subject to a 50% greater self-
discharge rate, a limited service life, and higher maintenance, and it is more expensive than the NiCad battery.

Directive 2006/66/EC of the European Union prohibits the placing on the market of portable batteries that contain more than 0.002% of
cadmium by weight. The aim of this directive was to improve "the environmental performance of batteries and accumulators"

Lead–Acid (Lead Storage) Battery


The lead–acid battery is used to provide the starting power in virtually every automobile and marine engine on the market. Marine and car batteries
typically consist of multiple cells connected in series. The total voltage generated by the battery is the potential per cell (E°cell) times the number of
cells.

2.6.4 https://chem.libretexts.org/@go/page/48482
Figure 2.6.3 : One Cell of a Lead–Acid Battery. The anodes in each cell of a rechargeable battery are plates or grids of lead containing spongy lead
metal, while the cathodes are similar grids containing powdered lead dioxide (PbO2). The electrolyte is an aqueous solution of sulfuric acid. The
value of E° for such a cell is about 2 V. Connecting three such cells in series produces a 6 V battery, whereas a typical 12 V car battery contains six
cells in series. When treated properly, this type of high-capacity battery can be discharged and recharged many times over.
As shown in Figure 2.6.3, the anode of each cell in a lead storage battery is a plate or grid of spongy lead metal, and the cathode is a similar grid
containing powdered lead dioxide (P bO ). The electrolyte is usually an approximately 37% solution (by mass) of sulfuric acid in water, with a
2

density of 1.28 g/mL (about 4.5 M H SO ). Because the redox active species are solids, there is no need to separate the electrodes. The electrode
2 4

reactions in each cell during discharge are as follows:


cathode (reduction):
\[PbO_{2(s)} + HSO^−_{4(aq)} + 3H^+_{(aq)} + 2e^− \rightarrow PbSO_{4(s)} + 2H_2O_{(l)} \label{Eq17} \]
with E ∘
cathode
= 1.685 V

anode (oxidation):
− + −
P b(s) + H S O → P bS O4(s) + H + 2e (2.6.8)
4(aq) (aq)

with E ∘
anode
= −0.356 V

overall:
− +
P b(s) + P b O2(s) + 2H S O + 2H → 2P bS O4(s) + 2 H2 O(l) (2.6.9)
4(aq) (aq)

and E ∘
cell
= 2.041 V

As the cell is discharged, a powder of P bSO forms on the electrodes. Moreover, sulfuric acid is consumed and water is produced, decreasing the
4

density of the electrolyte and providing a convenient way of monitoring the status of a battery by simply measuring the density of the electrolyte.
This is often done with the use of a hydrometer.

2.6.5 https://chem.libretexts.org/@go/page/48482
I'm not a robot
reCAPTCHA
Privacy - Terms

About this page

Our systems have detected unusual traffic from your computer


network. This page checks to see if it's really you sending the
requests, and not a robot. Why did this happen?

IP address: 2604:a880:2:d0::2056:8001
Time: 2023-03-12T20:39:54Z
URL: https://www.youtube-nocookie.com/embed/SRcOqfL6GqQ?
vq=hd1080

A hydrometer can be used to test the specific gravity of each cell as a measure of its state of charge (www.youtube.com/watch?v=SRcOqfL6GqQ).
When an external voltage in excess of 2.04 V per cell is applied to a lead–acid battery, the electrode reactions reverse, and P bSO is converted back
4

to metallic lead and P bO . If the battery is recharged too vigorously, however, electrolysis of water can occur:
2

2 H2 O(l) → 2 H2(g) + O2(g) (2.6.10)

This results in the evolution of potentially explosive hydrogen gas. The gas bubbles formed in this way can dislodge some of the P bSO or P bO4 2

particles from the grids, allowing them to fall to the bottom of the cell, where they can build up and cause an internal short circuit. Thus the
recharging process must be carefully monitored to optimize the life of the battery. With proper care, however, a lead–acid battery can be discharged
and recharged thousands of times. In automobiles, the alternator supplies the electric current that causes the discharge reaction to reverse.

Fuel Cells
A fuel cell is a galvanic cell that requires a constant external supply of reactants because the products of the reaction are continuously removed.
Unlike a battery, it does not store chemical or electrical energy; a fuel cell allows electrical energy to be extracted directly from a chemical reaction.
In principle, this should be a more efficient process than, for example, burning the fuel to drive an internal combustion engine that turns a generator,
which is typically less than 40% efficient, and in fact, the efficiency of a fuel cell is generally between 40% and 60%. Unfortunately, significant cost
and reliability problems have hindered the wide-scale adoption of fuel cells. In practice, their use has been restricted to applications in which mass
may be a significant cost factor, such as US manned space vehicles.

2.6.6 https://chem.libretexts.org/@go/page/48482
Figure 2.6.4 : A Hydrogen Fuel Cell Produces Electrical Energy Directly from a Chemical Reaction. Hydrogen is oxidized to protons at the anode,
and the electrons are transferred through an external circuit to the cathode, where oxygen is reduced and combines with H to form water. A solid
+

electrolyte allows the protons to diffuse from the anode to the cathode. Although fuel cells are an essentially pollution-free means of obtaining
electrical energy, their expense and technological complexity have thus far limited their applications.
These space vehicles use a hydrogen/oxygen fuel cell that requires a continuous input of H2(g) and O2(g), as illustrated in Figure 2.6.4 . The
electrode reactions are as follows:
cathode (reduction):
+ −
O2(g) + 4 H + 4e → 2 H2 O(g) (2.6.11)

anode (oxidation):
+ −
2 H2(g) → 4 H + 4e (2.6.12)

overall:
2 H2(g) + O2(g) → 2 H2 O(g) (2.6.13)

The overall reaction represents an essentially pollution-free conversion of hydrogen and oxygen to water, which in space vehicles is then collected
and used. Although this type of fuel cell should produce 1.23 V under standard conditions, in practice the device achieves only about 0.9 V. One of
the major barriers to achieving greater efficiency is the fact that the four-electron reduction of O (g) at the cathode is intrinsically rather slow, which
2

limits current that can be achieved. All major automobile manufacturers have major research programs involving fuel cells: one of the most
important goals is the development of a better catalyst for the reduction of O (g) .
2

Summary
Commercial batteries are galvanic cells that use solids or pastes as reactants to maximize the electrical output per unit mass. A battery is a contained
unit that produces electricity, whereas a fuel cell is a galvanic cell that requires a constant external supply of one or more reactants to generate
electricity. One type of battery is the Leclanché dry cell, which contains an electrolyte in an acidic water-based paste. This battery is called an
alkaline battery when adapted to operate under alkaline conditions. Button batteries have a high output-to-mass ratio; lithium–iodine batteries consist
of a solid electrolyte; the nickel–cadmium (NiCad) battery is rechargeable; and the lead–acid battery, which is also rechargeable, does not require the
electrodes to be in separate compartments. A fuel cell requires an external supply of reactants as the products of the reaction are continuously
removed. In a fuel cell, energy is not stored; electrical energy is provided by a chemical reaction.

2.6: Batteries- Producing Electricity Through Chemical Reactions is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.
20.7: Batteries and Fuel Cells is licensed CC BY-NC-SA 3.0.

2.6.7 https://chem.libretexts.org/@go/page/48482
2.7: Corrosion- Unwanted Voltaic Cells
 Learning Objectives
To understand the process of corrosion.

Corrosion is a galvanic process by which metals deteriorate through oxidation—usually but not always to their oxides. For
example, when exposed to air, iron rusts, silver tarnishes, and copper and brass acquire a bluish-green surface called a patina. Of
the various metals subject to corrosion, iron is by far the most important commercially. An estimated $100 billion per year is spent
in the United States alone to replace iron-containing objects destroyed by corrosion. Consequently, the development of methods for
protecting metal surfaces from corrosion constitutes a very active area of industrial research. In this section, we describe some of
the chemical and electrochemical processes responsible for corrosion. We also examine the chemical basis for some common
methods for preventing corrosion and treating corroded metals.

Corrosion is a REDOX process.


Under ambient conditions, the oxidation of most metals is thermodynamically spontaneous, with the notable exception of gold and
platinum. Hence it is actually somewhat surprising that any metals are useful at all in Earth’s moist, oxygen-rich atmosphere. Some
metals, however, are resistant to corrosion for kinetic reasons. For example, aluminum in soft-drink cans and airplanes is protected
by a thin coating of metal oxide that forms on the surface of the metal and acts as an impenetrable barrier that prevents further
destruction. Aluminum cans also have a thin plastic layer to prevent reaction of the oxide with acid in the soft drink. Chromium,
magnesium, and nickel also form protective oxide films. Stainless steels are remarkably resistant to corrosion because they usually
contain a significant proportion of chromium, nickel, or both.
In contrast to these metals, when iron corrodes, it forms a red-brown hydrated metal oxide (Fe O ⋅ xH O ), commonly known as
2 3 2

rust, that does not provide a tight protective film (Figure 2.7.1). Instead, the rust continually flakes off to expose a fresh metal
surface vulnerable to reaction with oxygen and water. Because both oxygen and water are required for rust to form, an iron nail
immersed in deoxygenated water will not rust—even over a period of several weeks. Similarly, a nail immersed in an organic
solvent such as kerosene or mineral oil will not rust because of the absence of water even if the solvent is saturated with oxygen.

Figure 2.7.1 : Rust, the Result of Corrosion of Metallic Iron. Iron is oxidized to Fe2+(aq) at an anodic site on the surface of the iron,
which is often an impurity or a lattice defect. Oxygen is reduced to water at a different site on the surface of the iron, which acts as
the cathode. Electrons are transferred from the anode to the cathode through the electrically conductive metal. Water is a solvent for
the Fe2+ that is produced initially and acts as a salt bridge. Rust (Fe2O3•xH2O) is formed by the subsequent oxidation of Fe2+ by
atmospheric oxygen. (CC BY-NC-SA; anonymous)
In the corrosion process, iron metal acts as the anode in a galvanic cell and is oxidized to Fe2+; oxygen is reduced to water at the
cathode. The relevant reactions are as follows:
at cathode:
+ −
O (g) + 4 H (aq) + 4 e ⟶ 2 H O(l)
2 2

with E o
SRP
= 1.23 V .

2.7.1 https://chem.libretexts.org/@go/page/48483
at anode:
2 + −
Fe(s) ⟶ Fe (aq) + 2 e

with E o
SRP
= −0.45 V .
overall:
+ 2 +
2 Fe(s) + O (g) + 4 H (aq) ⟶ 2 Fe (aq) + 2 H O(l) (2.7.1)
2 2

with E o
cell
= 1.68 V .
The Fe 2 +
ions produced in the initial reaction are then oxidized by atmospheric oxygen to produce the insoluble hydrated oxide
containing Fe , as represented in the following equation:
3 +

2 + +
4 Fe (aq) + O (g) + (2 + 4 x)H O → 2 Fe O ⋅ xH O + 4 H (aq) (2.7.2)
2 2 2 3 2

The sign and magnitude of E for the corrosion process (Equation 2.7.1) indicate that there is a strong driving force for the
o
cell

oxidation of iron by O2 under standard conditions (1 M H+). Under neutral conditions, the driving force is somewhat less but still
appreciable (E = 1.25 V at pH 7.0). Normally, the reaction of atmospheric CO2 with water to form H+ and HCO3− provides a low
enough pH to enhance the reaction rate, as does acid rain. Automobile manufacturers spend a great deal of time and money
developing paints that adhere tightly to the car’s metal surface to prevent oxygenated water, acid, and salt from coming into contact
with the underlying metal. Unfortunately, even the best paint is subject to scratching or denting, and the electrochemical nature of
the corrosion process means that two scratches relatively remote from each other can operate together as anode and cathode,
leading to sudden mechanical failure (Figure 2.7.2).

Figure 2.7.2 : Small Scratches in a Protective Paint Coating Can Lead to the Rapid Corrosion of Iron. Holes in a protective coating
allow oxygen to be reduced at the surface with the greater exposure to air (the cathode), while metallic iron is oxidized to Fe2+(aq)
at the less exposed site (the anode). Rust is formed when Fe2+(aq) diffuses to a location where it can react with atmospheric
oxygen, which is often remote from the anode. The electrochemical interaction between cathodic and anodic sites can cause a large
pit to form under a painted surface, eventually resulting in sudden failure with little visible warning that corrosion has occurred.

Prophylactic Protection
One of the most common techniques used to prevent the corrosion of iron is applying a protective coating of another metal that is
more difficult to oxidize. Faucets and some external parts of automobiles, for example, are often coated with a thin layer of
chromium using an electrolytic process. With the increased use of polymeric materials in cars, however, the use of chrome-plated
steel has diminished in recent years. Similarly, the “tin cans” that hold soups and other foods are actually consist of steel container
that is coated with a thin layer of tin. While neither chromium nor tin metals are intrinsically resistant to corrosion, they both form
protective oxide coatings that hinder access of oxygen and water to the underlying steel (iron alloy).

2.7.2 https://chem.libretexts.org/@go/page/48483
Figure 2.7.3 : Galvanic Corrosion. If iron is in contact with a more corrosion-resistant metal such as tin, copper, or lead, the other
metal can act as a large cathode that greatly increases the rate of reduction of oxygen. Because the reduction of oxygen is coupled
to the oxidation of iron, this can result in a dramatic increase in the rate at which iron is oxidized at the anode. Galvanic corrosion
is likely to occur whenever two dissimilar metals are connected directly, allowing electrons to be transferred from one to the other.
As with a protective paint, scratching a protective metal coating will allow corrosion to occur. In this case, however, the presence of
the second metal can actually increase the rate of corrosion. The values of the standard electrode potentials for Sn (E° = −0.14
2 +

V) and Fe2+ (E° = −0.45 V) in Table P2 show that Fe is more easily oxidized than Sn. As a result, the more corrosion-resistant
metal (in this case, tin) accelerates the corrosion of iron by acting as the cathode and providing a large surface area for the
reduction of oxygen (Figure 2.7.3). This process is seen in some older homes where copper and iron pipes have been directly
connected to each other. The less easily oxidized copper acts as the cathode, causing iron to dissolve rapidly near the connection
and occasionally resulting in a catastrophic plumbing failure.

Cathodic Protection
One way to avoid these problems is to use a more easily oxidized metal to protect iron from corrosion. In this approach, called
cathodic protection, a more reactive metal such as Zn (E° = −0.76 V for Zn + 2 e ⟶ Zn ) becomes the anode, and iron
2 + −

becomes the cathode. This prevents oxidation of the iron and protects the iron object from corrosion. The reactions that occur under
these conditions are as follows:
− +
O2(g) + 4 e + 4H → 2 H2 O(l) (2.7.3)
(aq)

reduction at cathode

2+ −
Z n(s) → Z n + 2e (2.7.4)
(aq)

oxidation at anode

+ 2+
2Z n(s) + O2(g) + 4 H → 2Z n + 2 H2 O(l) (2.7.5)
(aq) (aq)

overall

The more reactive metal reacts with oxygen and will eventually dissolve, “sacrificing” itself to protect the iron object. Cathodic
protection is the principle underlying galvanized steel, which is steel protected by a thin layer of zinc. Galvanized steel is used in
objects ranging from nails to garbage cans.

Crystalline surface of a hot-dip galvanized steel surface. This served both as prophylactic protection (protecting the underlying
steel from the oxygen in the air) and cathodic protection (once exposed, the zinc will oxidize before the underlying steel).

2.7.3 https://chem.libretexts.org/@go/page/48483
In a similar strategy, sacrificial electrodes using magnesium, for example, are used to protect underground tanks or pipes (Figure
2.7.4). Replacing the sacrificial electrodes is more cost-effective than replacing the iron objects they are protecting.

Figure 2.7.4 : The Use of a Sacrificial Electrode to Protect Against Corrosion. Connecting a magnesium rod to an underground steel
pipeline protects the pipeline from corrosion. Because magnesium (E° = −2.37 V) is much more easily oxidized than iron (E° =
−0.45 V), the Mg rod acts as the anode in a galvanic cell. The pipeline is therefore forced to act as the cathode at which oxygen is
reduced. The soil between the anode and the cathode acts as a salt bridge that completes the electrical circuit and maintains
electrical neutrality. As Mg(s) is oxidized to Mg2+ at the anode, anions in the soil, such as nitrate, diffuse toward the anode to
neutralize the positive charge. Simultaneously, cations in the soil, such as H+ or NH4+, diffuse toward the cathode, where they
replenish the protons that are consumed as oxygen is reduced. A similar strategy uses many miles of somewhat less reactive zinc
wire to protect the Alaska oil pipeline.

 Example 2.7.1

Suppose an old wooden sailboat, held together with iron screws, has a bronze propeller (recall that bronze is an alloy of copper
containing about 7%–10% tin).
a. If the boat is immersed in seawater, what corrosion reaction will occur? What is E o
°cell ?
b. How could you prevent this corrosion from occurring?
Given: identity of metals
Asked for: corrosion reaction, E o
°cell , and preventive measures

Strategy:
A. Write the reactions that occur at the anode and the cathode. From these, write the overall cell reaction and calculate E o
°cell .
B. Based on the relative redox activity of various substances, suggest possible preventive measures.

Solution
a. A According to Table P2, both copper and tin are less active metals than iron (i.e., they have higher positive values of
o
E ° cellthan iron). Thus if tin or copper is brought into electrical contact by seawater with iron in the presence of oxygen,
corrosion will occur. We therefore anticipate that the bronze propeller will act as the cathode at which O is reduced, and 2

the iron screws will act as anodes at which iron dissolves:


+ − ∘
cathode: O2 (s) + 4 H (aq) + 4 e → 2 H2 O(l) E = 1.23 V
cathode

2+ − ∘
anode: Fe(s) → F e + 2e E = −0.45 V
anode

+ 2+ ∘
overall: 2Fe(s) + O2 (g) + 4 H (aq) → 2F e (aq) + 2 H2 O(l) E = 1.68 V
overall

Over time, the iron screws will dissolve, and the boat will fall apart.
b. B Possible ways to prevent corrosion, in order of decreasing cost and inconvenience, are as follows: disassembling the boat
and rebuilding it with bronze screws; removing the boat from the water and storing it in a dry place; or attaching an
inexpensive piece of zinc metal to the propeller shaft to act as a sacrificial electrode and replacing it once or twice a year.

2.7.4 https://chem.libretexts.org/@go/page/48483
Because zinc is a more active metal than iron, it will act as the sacrificial anode in the electrochemical cell and dissolve
(Equation 2.7.5).

Zinc sacrificial anode (rounded object screwed to underside of hull) used to prevent corrosion on the screw in a boat via
cathodic protection. Image by Rémi Kaupp and used with permission.

 Exercise 2.7.1

Suppose the water pipes leading into your house are made of lead, while the rest of the plumbing in your house is iron. To
eliminate the possibility of lead poisoning, you call a plumber to replace the lead pipes. He quotes you a very low price if he
can use up his existing supply of copper pipe to do the job.
a. Do you accept his proposal?
b. What else should you have the plumber do while at your home?

Answer a
Not unless you plan to sell the house very soon because the Cu/Fe pipe joints will lead to rapid corrosion.
Answer b
Any existing Pb/Fe joints should be examined carefully for corrosion of the iron pipes due to the Pb– Fe junction; the
less active Pb will have served as the cathode for the reduction of O , promoting oxidation of the more active Fe nearby.
2

Summary
Corrosion is a galvanic process that can be prevented using cathodic protection. The deterioration of metals through oxidation is a
galvanic process called corrosion. Protective coatings consist of a second metal that is more difficult to oxidize than the metal
being protected. Alternatively, a more easily oxidized metal can be applied to a metal surface, thus providing cathodic protection of
the surface. A thin layer of zinc protects galvanized steel. Sacrificial electrodes can also be attached to an object to protect it.

2.7: Corrosion- Unwanted Voltaic Cells is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
20.8: Corrosion is licensed CC BY-NC-SA 3.0.

2.7.5 https://chem.libretexts.org/@go/page/48483
2.8: Electrolysis- Causing Nonspontaneous Reactions to Occur
 Learning Objectives
To understand electrolysis and describe it quantitatively.

In this chapter, we have described various galvanic cells in which a spontaneous chemical reaction is used to generate electrical
energy. In an electrolytic cell, however, the opposite process, called electrolysis, occurs: an external voltage is applied to drive a
nonspontaneous reaction. In this section, we look at how electrolytic cells are constructed and explore some of their many
commercial applications.

Electrolytic Cells
If we construct an electrochemical cell in which one electrode is copper metal immersed in a 1 M Cu2+ solution and the other
electrode is cadmium metal immersed in a 1 M C d solution and then close the circuit, the potential difference between the two
2+

compartments will be 0.74 V. The cadmium electrode will begin to dissolve (Cd is oxidized to Cd2+) and is the anode, while
metallic copper will be deposited on the copper electrode (Cu2+ is reduced to Cu), which is the cathode (Figure 2.8.1a).

Figure 2.8.1 : An Applied Voltage Can Reverse the Flow of Electrons in a Galvanic Cd/Cu Cell. (a) When compartments that
contain a Cd electrode immersed in 1 M Cd2+(aq) and a Cu electrode immersed in 1 M Cu2+(aq) are connected to create a galvanic
cell, Cd(s) is spontaneously oxidized to Cd2+(aq) at the anode, and Cu2+(aq) is spontaneously reduced to Cu(s) at the cathode. The
potential of the galvanic cell is 0.74 V. (b) Applying an external potential greater than 0.74 V in the reverse direction forces
electrons to flow from the Cu electrode [which is now the anode, at which metallic Cu(s) is oxidized to Cu2+(aq)] and into the Cd
electrode [which is now the cathode, at which Cd2+(aq) is reduced to Cd(s)]. The anode in an electrolytic cell is positive because
electrons are flowing from it, whereas the cathode is negative because electrons are flowing into it. (CC BY-SA-NC; anonymous)
The overall reaction is as follows:
2 + 2 +
Cd(s) + Cu (aq) → Cd (aq) + Cu(s)

with E °cell = 0.74 V

This reaction is thermodynamically spontaneous as written (ΔG o


<0 ):
∘ ∘
ΔG = −nF E
cell


= −(2 mol e )[96, 485 J/(V ⋅ mol)](0.74 V)

= −140 kJ (per mole Cd)

2.8.1 https://chem.libretexts.org/@go/page/48484
In this direction, the system is acting as a galvanic cell.

In an electrolytic cell, an external voltage is applied to drive a nonspontaneous reaction.


The reverse reaction, the reduction of Cd2+ by Cu, is thermodynamically nonspontaneous and will occur only with an input of 140
kJ. We can force the reaction to proceed in the reverse direction by applying an electrical potential greater than 0.74 V from an
external power supply. The applied voltage forces electrons through the circuit in the reverse direction, converting a galvanic cell to
an electrolytic cell. Thus the copper electrode is now the anode (Cu is oxidized), and the cadmium electrode is now the cathode
(Cd2+ is reduced) (Figure 2.8.1b). The signs of the cathode and the anode have switched to reflect the flow of electrons in the
circuit. The half-reactions that occur at the cathode and the anode are as follows:
half-reaction at the cathode:
2 + −
Cd (aq) + 2 e → Cd(s) (2.8.1)

with E ∘
cathode
= −0.40 V

half-reaction at the anode:


2 + −
Cu(s) → Cu (aq) + 2 e (2.8.2)

with E ∘
anode
= 0.34 V

Overall Reaction:
2 + 2 +
Cd (aq) + Cu(s) → Cd(s) + Cu (aq) (2.8.3)

with E ∘
cell
= −0.74 V

Because E < 0 , the overall reaction—the reduction of C d by C u—clearly cannot occur spontaneously and proceeds only

cell
2+

when sufficient electrical energy is applied. The differences between galvanic and electrolytic cells are summarized in Table 2.8.1.
Table 2.8.1 : Comparison of Galvanic and Electrolytic Cells
Property Galvanic Cell Electrolytic Cell

ΔG <0 >0

Ecell >0 <0

Electrode Process

anode oxidation oxidation

cathode reduction reduction

Sign of Electrode

anode − +

cathode + −

Electrolytic Reactions
At sufficiently high temperatures, ionic solids melt to form liquids that conduct electricity extremely well due to the high
concentrations of ions. If two inert electrodes are inserted into molten NaCl, for example, and an electrical potential is applied,
Cl

is oxidized at the anode, and Na is reduced at the cathode. The overall reaction is as follows:
+

2 NaCl(l) → 2 Na(l) + Cl (g) (2.8.4)


2

This is the reverse of the formation of NaCl from its elements. The product of the reduction reaction is liquid sodium because the
melting point of sodium metal is 97.8°C, well below that of NaCl (801°C). Approximately 20,000 tons of sodium metal are
produced commercially in the United States each year by the electrolysis of molten NaCl in a Downs cell (Figure 2.8.2). In this
specialized cell, CaCl (melting point = 772°C) is first added to the NaCl to lower the melting point of the mixture to about
2

600°C, thereby lowering operating costs.

2.8.2 https://chem.libretexts.org/@go/page/48484
Figure 2.8.2 : A Downs Cell for the Electrolysis of Molten NaCl. The electrolysis of a molten mixture of NaCl and CaCl2 results in
the formation of elemental sodium and chlorine gas. Because sodium is a liquid under these conditions and liquid sodium is less
dense than molten sodium chloride, the sodium floats to the top of the melt and is collected in concentric capped iron cylinders
surrounding the cathode. Gaseous chlorine collects in the inverted cone over the anode. An iron screen separating the cathode and
anode compartments ensures that the molten sodium and gaseous chlorine do not come into contact. (CC BY-SA-NC; anonymous)

Similarly, in the Hall–Heroult process used to produce aluminum commercially, a molten mixture of about 5% aluminum oxide
(Al2O3; melting point = 2054°C) and 95% cryolite (Na3AlF6; melting point = 1012°C) is electrolyzed at about 1000°C, producing
molten aluminum at the cathode and CO2 gas at the carbon anode. The overall reaction is as follows:
2 Al O (l) + 3 C(s) ⟶ 4 Al(l) + 3 CO (g) (2.8.5)
2 3 2

Oxide ions react with oxidized carbon at the anode, producing CO2(g).
There are two important points to make about these two commercial processes and about the electrolysis of molten salts in general.
1. The electrode potentials for molten salts are likely to be very different from the standard cell potentials listed in Table P2, which
are compiled for the reduction of the hydrated ions in aqueous solutions under standard conditions.
2. Using a mixed salt system means there is a possibility of competition between different electrolytic reactions. When a mixture
of NaCl and CaCl2 is electrolyzed, Cl− is oxidized because it is the only anion present, but either Na+ or Ca2+ can be reduced.
Conversely, in the Hall–Heroult process, only one cation is present that can be reduced (Al3+), but there are three species that
can be oxidized: C, O2−, and F−.
In the Hall–Heroult process, C is oxidized instead of O2− or F− because oxygen and fluorine are more electronegative than carbon,
which means that C is a weaker oxidant than either O2 or F2. Similarly, in the Downs cell, we might expect electrolysis of a
NaCl/CaCl2 mixture to produce calcium rather than sodium because Na is slightly less electronegative than Ca (χ = 0.93 versus
1.00, respectively), making Na easier to oxidize and, conversely, Na+ more difficult to reduce. In fact, the reduction of Na+ to Na is
the observed reaction. In cases where the electronegativities of two species are similar, other factors, such as the formation of
complex ions, become important and may determine the outcome.

 Example 2.8.1

If a molten mixture of MgCl2 and KBr is electrolyzed, what products will form at the cathode and the anode, respectively?
Given: identity of salts
Asked for: electrolysis products

2.8.3 https://chem.libretexts.org/@go/page/48484
Strategy:
A. List all the possible reduction and oxidation products. Based on the electronegativity values shown in Figure 7.5, determine
which species will be reduced and which species will be oxidized.
B. Identify the products that will form at each electrode.

Solution
A The possible reduction products are Mg and K, and the possible oxidation products are Cl2 and Br2. Because Mg is more
electronegative than K (χ = 1.31 versus 0.82), it is likely that Mg will be reduced rather than K. Because Cl is more
electronegative than Br (3.16 versus 2.96), Cl2 is a stronger oxidant than Br2.
B Electrolysis will therefore produce Br2 at the anode and Mg at the cathode.

 Exercise 2.8.1

Predict the products if a molten mixture of AlBr3 and LiF is electrolyzed.

Answer
Br2 and Al

Electrolysis can also be used to drive the thermodynamically nonspontaneous decomposition of water into its constituent elements:
H2 and O2. However, because pure water is a very poor electrical conductor, a small amount of an ionic solute (such as H2SO4 or
Na2SO4) must first be added to increase its electrical conductivity. Inserting inert electrodes into the solution and applying a
voltage between them will result in the rapid evolution of bubbles of H2 and O2 (Figure 2.8.3).

Figure 2.8.3 : The Electrolysis of Water. Applying an external potential of about 1.7–1.9 V to two inert electrodes immersed in an
aqueous solution of an electrolyte such as H2SO4 or Na2SO4 drives the thermodynamically nonspontaneous decomposition of water
into H2 at the cathode and O2 at the anode. (CC BY-SA-NC; anonymous)
The reactions that occur are as follows:
cathode:
+ − ∘
2H + 2e → H2(g)   E = 0V (2.8.6)
(aq) cathode

anode:
+ − ∘
2 H2 O(l) → O2(g) + 4 H + 4e  E = 1.23 V (2.8.7)
(aq) anode

overall:

2 H2 O(l) → O2(g) + 2 H2(g)  E = −1.23 V (2.8.8)
cell

2.8.4 https://chem.libretexts.org/@go/page/48484
For a system that contains an electrolyte such as Na2SO4, which has a negligible effect on the ionization equilibrium of liquid
water, the pH of the solution will be 7.00 and [H+] = [OH−] = 1.0 × 10−7. Assuming that P = P = 1 atm, we can use the
O2 H2

standard potentials to calculate E for the overall reaction:


0.0591 V
∘ 2
Ecell = E −( ) log(PO2 P ) (2.8.9)
cell H2
n

0.0591 V
= −1.23 V − ( ) log(1) = −1.23 V (2.8.10)
4

Thus Ecell is −1.23 V, which is the value of E°cell if the reaction is carried out in the presence of 1 M H+ rather than at pH 7.0.
In practice, a voltage about 0.4–0.6 V greater than the calculated value is needed to electrolyze water. This added voltage, called an
overvoltage, represents the additional driving force required to overcome barriers such as the large activation energy for the
formation of a gas at a metal surface. Overvoltages are needed in all electrolytic processes, which explain why, for example,
approximately 14 V must be applied to recharge the 12 V battery in your car.
In general, any metal that does not react readily with water to produce hydrogen can be produced by the electrolytic reduction of an
aqueous solution that contains the metal cation. The p-block metals and most of the transition metals are in this category, but metals
in high oxidation states, which form oxoanions, cannot be reduced to the metal by simple electrolysis. Active metals, such as
aluminum and those of groups 1 and 2, react so readily with water that they can be prepared only by the electrolysis of molten salts.
Similarly, any nonmetallic element that does not readily oxidize water to O2 can be prepared by the electrolytic oxidation of an
aqueous solution that contains an appropriate anion. In practice, among the nonmetals, only F2 cannot be prepared using this
method. Oxoanions of nonmetals in their highest oxidation states, such as NO3−, SO42−, PO43−, are usually difficult to reduce
electrochemically and usually behave like spectator ions that remain in solution during electrolysis.

In general, any metal that does not react readily with water to produce hydrogen can be
produced by the electrolytic reduction of an aqueous solution that contains the metal
cation.

Electroplating
In a process called electroplating, a layer of a second metal is deposited on the metal electrode that acts as the cathode during
electrolysis. Electroplating is used to enhance the appearance of metal objects and protect them from corrosion. Examples of
electroplating include the chromium layer found on many bathroom fixtures or (in earlier days) on the bumpers and hubcaps of
cars, as well as the thin layer of precious metal that coats silver-plated dinnerware or jewelry. In all cases, the basic concept is the
same. A schematic view of an apparatus for electroplating silverware and a photograph of a commercial electroplating cell are
shown in Figure 2.8.4.

2.8.5 https://chem.libretexts.org/@go/page/48484
Figure 2.8.3 : Electroplating. (a) Electroplating uses an electrolytic cell in which the object to be plated, such as a fork, is immersed
in a solution of the metal to be deposited. The object being plated acts as the cathode, on which the desired metal is deposited in a
thin layer, while the anode usually consists of the metal that is being deposited (in this case, silver) that maintains the solution
concentration as it dissolves. (b) In this commercial electroplating apparatus, a large number of objects can be plated
simultaneously by lowering the rack into the Ag solution and applying the correct potential. (CC BY-SA-NC; anonymous)
+

The half-reactions in electroplating a fork, for example, with silver are as follows:
cathode (fork):
+ −
Ag (aq) + e ⟶ Ag(s)   E °cathode = 0.80V  

anode (silver bar):


+ −
Ag(s) ⟶ Ag (aq) + e   E °anode = 0.80V

The overall reaction is the transfer of silver metal from one electrode (a silver bar acting as the anode) to another (a fork acting as
the cathode). Because E = 0 V , it takes only a small applied voltage to drive the electroplating process. In practice, various
o
cell

other substances may be added to the plating solution to control its electrical conductivity and regulate the concentration of free
metal ions, thus ensuring a smooth, even coating.

Quantitative Considerations
If we know the stoichiometry of an electrolysis reaction, the amount of current passed, and the length of time, we can calculate the
amount of material consumed or produced in a reaction. Conversely, we can use stoichiometry to determine the combination of
current and time needed to produce a given amount of material.
The quantity of material that is oxidized or reduced at an electrode during an electrochemical reaction is determined by the
stoichiometry of the reaction and the amount of charge that is transferred. For example, in the reaction
+ −
Ag (aq) + e → Ag(s)

1 mol of electrons reduces 1 mol of Ag to Ag metal. In contrast, in the reaction


+

2 + −
Cu (aq) + 2 e → Cu(s)

1 mol of electrons reduces only 0.5 mol of Cu to Cu metal. Recall that the charge on 1 mol of electrons is 1 faraday (1 F),
2 +

which is equal to 96,485 C. We can therefore calculate the number of moles of electrons transferred when a known current is
passed through a cell for a given period of time. The total charge (q in coulombs) transferred is the product of the current (I in
amperes) and the time (t , in seconds):

q = I ×t (2.8.11)

2.8.6 https://chem.libretexts.org/@go/page/48484
The stoichiometry of the reaction and the total charge transferred enable us to calculate the amount of product formed during an
electrolysis reaction or the amount of metal deposited in an electroplating process.
For example, if a current of 0.60 A passes through an aqueous solution of CuSO
4
for 6.0 min, the total number of coulombs of
charge that passes through the cell is as follows:

q = (0.60 A)(6.0 min)(60 s/min)

= 220 A ⋅ s

= 220 C

The number of moles of electrons transferred to Cu 2 +


is therefore


220 C
moles e =
96,485 C/mol

−3 −
= 2.3 × 10  mol e

Because two electrons are required to reduce a single Cu2+ ion, the total number of moles of Cu produced is half the number of
moles of electrons transferred, or 1.2 × 10−3 mol. This corresponds to 76 mg of Cu. In commercial electrorefining processes, much
higher currents (greater than or equal to 50,000 A) are used, corresponding to approximately 0.5 F/s, and reaction times are on the
order of 3–4 weeks.

 Example 2.8.2

A silver-plated spoon typically contains about 2.00 g of Ag. If 12.0 h are required to achieve the desired thickness of the Ag
coating, what is the average current per spoon that must flow during the electroplating process, assuming an efficiency of
100%?
Given: mass of metal, time, and efficiency
Asked for: current required

Strategy:
A. Calculate the number of moles of metal corresponding to the given mass transferred.
B. Write the reaction and determine the number of moles of electrons required for the electroplating process.
C. Use the definition of the faraday to calculate the number of coulombs required. Then convert coulombs to current in
amperes.

Solution
A We must first determine the number of moles of Ag corresponding to 2.00 g of Ag:
2.00 g −2
moles Ag = = 1.85 × 10  mol Ag
107.868 g/mol

B The reduction reaction is Ag+(aq) + e− → Ag(s), so 1 mol of electrons produces 1 mol of silver.
C Using the definition of the faraday,
coulombs = (1.85 × 10−2mol e−)(96,485 C/mol e−) = 1.78 × 103 C / mole
The current in amperes needed to deliver this amount of charge in 12.0 h is therefore
3
1.78 × 10  C
amperes =
(12.0 h)(60 min/h)(60 s/min)
−2 −2
= 4.12 × 10  C/s = 4.12 × 10  A

Because the electroplating process is usually much less than 100% efficient (typical values are closer to 30%), the actual
current necessary is greater than 0.1 A.

2.8.7 https://chem.libretexts.org/@go/page/48484
 Exercise 2.8.2
A typical aluminum soft-drink can weighs about 29 g. How much time is needed to produce this amount of Al(s) in the Hall–
Heroult process, using a current of 15 A to reduce a molten Al2O3/Na3AlF6 mixture?

Answer
5.8 h

I'm not a robot


reCAPTCHA
Privacy - Terms

About this page

Our systems have detected unusual traffic from your


computer network This page checks to see if it's really you
Electroplating: Electroplating(opens in new window) [youtu.be]

Summary
In electrolysis, an external voltage is applied to drive a nonspontaneous reaction. The quantity of material oxidized or reduced can
be calculated from the stoichiometry of the reaction and the amount of charge transferred. Relationship of charge, current and time:

q = I ×t

In electrolysis, an external voltage is applied to drive a nonspontaneous reaction. Electrolysis can also be used to produce H2 and
O2 from water. In practice, an additional voltage, called an overvoltage, must be applied to overcome factors such as a large
activation energy and a junction potential. Electroplating is the process by which a second metal is deposited on a metal surface,
thereby enhancing an object’s appearance or providing protection from corrosion. The amount of material consumed or produced in
a reaction can be calculated from the stoichiometry of an electrolysis reaction, the amount of current passed, and the duration of the
electrolytic reaction.

2.8: Electrolysis- Causing Nonspontaneous Reactions to Occur is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by LibreTexts.
20.9: Electrolysis is licensed CC BY-NC-SA 3.0.

2.8.8 https://chem.libretexts.org/@go/page/48484
CHAPTER OVERVIEW
3: Coordination Chemistry
Complexes or coordination compounds are molecules that posess a metal center that is bound to ligands (atoms, ions, or
molecules that donate electrons to the metal). These complexes can be neutral or charged. When the complex is charged, it is
stabilized by neighboring counter-ions.
3.1: Werner’s Theory of Coordination Compounds
3.2: Ligands
3.3: Nomenclature
3.4: Isomerism
3.5: Optical Isomerism in Metal Complexes
3.6: Stability Aspects of Complex-Ion Equilibria
3.7: Chelation
3.8: Bonding in Octahedral Complex Ions- Crystal Field Theory
3.9: Bonding in Non-Octahedral Complex Ions- Crystal Field Theory
3.10: Magnetic Behavior of Atoms, Molecules, and Materials
3.11: Magnetic Behavior of Complex Ions
3.12: Optical Properties of Coordination Compounds (Color)

3: Coordination Chemistry is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

1
3.1: Werner’s Theory of Coordination Compounds
 Learning Objectives
To know the most common structures observed for metal complexes.
To predict the relative stabilities of metal complexes with different ligands

One of the most important properties of metallic elements is their ability to act as Lewis acids that form complexes with a variety
of Lewis bases. A metal complex consists of a central metal atom or ion that is bonded to one or more ligands (from the Latin
ligare, meaning “to bind”), which are ions or molecules that contain one or more pairs of electrons that can be shared with the
metal. Metal complexes can be neutral, such as Co(NH3)3Cl3; positively charged, such as [Nd(H2O)9]3+; or negatively charged,
such as [UF8]4−. Electrically charged metal complexes are sometimes called complex ions. A coordination compound contains one
or more metal complexes.
Coordination compounds are important for at least three reasons. First, most of the elements in the periodic table are metals, and
almost all metals form complexes, so metal complexes are a feature of the chemistry of more than half the elements. Second, many
industrial catalysts are metal complexes, and such catalysts are steadily becoming more important as a way to control reactivity.
For example, a mixture of a titanium complex and an organometallic compound of aluminum is the catalyst used to produce most
of the polyethylene and polypropylene “plastic” items we use every day. Finally, transition-metal complexes are essential in
biochemistry. Examples include hemoglobin, an iron complex that transports oxygen in our blood; cytochromes, iron complexes
that transfer electrons in our cells; and complexes of Fe, Zn, Cu, and Mo that are crucial components of certain enzymes, the
catalysts for all biological reactions.

History of the Coordination Compounds


Coordination compounds have been known and used since antiquity; probably the oldest is the deep blue pigment called Prussian
blue: KFe (CN) . The chemical nature of these substances, however, was unclear for a number of reasons. For example, many
2 6

compounds called “double salts” were known, such as AlF ⋅ 3 KF , Fe(CN) ⋅ 4 KCN , and ZnCl ⋅ 2 CsCl , which were
3 2 2

combinations of simple salts in fixed and apparently arbitrary ratios. Why should AlF ⋅ 3 KF exist but not AlF ⋅ 4 KF or
3 3

AlF ⋅ 2 KF ? And why should a 3:1 KF:AlF3 mixture have different chemical and physical properties than either of its
3

components? Similarly, adducts of metal salts with neutral molecules such as ammonia were also known—for example,
CoCl ⋅ 6 NH , which was first prepared sometime before 1798. Like the double salts, the compositions of these adducts exhibited
3 3

fixed and apparently arbitrary ratios of the components. For example, CoCl ⋅ 6 NH , CoCl ⋅ 5 NH , CoCl ⋅ 4 NH , and
3 3 3 3 3 3

CoCl ⋅ 3 NH
3
were all known and had very different properties, but despite all attempts, chemists could not prepare
3

CoCl ⋅ 2 NH
3
or CoCl ⋅ NH .
3 3 3

Although the chemical composition of such compounds was readily established by existing analytical methods, their chemical
nature was puzzling and highly controversial. The major problem was that what we now call valence (i.e., the oxidation state) and
coordination number were thought to be identical. As a result, highly implausible (to modern eyes at least) structures were
proposed for such compounds, including the “Chattanooga choo-choo” model for CoCl3·4NH3 shown here.

The modern theory of coordination chemistry is based largely on the work of Alfred Werner (1866–1919; Nobel Prize in Chemistry
in 1913). In a series of careful experiments carried out in the late 1880s and early 1890s, he examined the properties of several
series of metal halide complexes with ammonia. For example, five different “adducts” of ammonia with PtCl4 were known at the
time: PtCl4·nNH3 (n = 2–6). Some of Werner’s original data on these compounds are shown in Table 3.1.1. The electrical
conductivity of aqueous solutions of these compounds was roughly proportional to the number of ions formed per mole, while the
number of chloride ions that could be precipitated as AgCl after adding Ag+(aq) was a measure of the number of “free” chloride
ions present. For example, Werner’s data on PtCl4·6NH3 in Table 3.1.1 showed that all the chloride ions were present as free
chloride. In contrast, PtCl4·2NH3 was a neutral molecule that contained no free chloride ions.

3.1.1 https://chem.libretexts.org/@go/page/48465
 Alfred Werner (1866–1919)

Werner, the son of a factory worker, was born in Alsace. He developed an interest in chemistry at an early age, and he did his
first independent research experiments at age 18. While doing his military service in southern Germany, he attended a series of
chemistry lectures, and he subsequently received his PhD at the University of Zurich in Switzerland, where he was appointed
professor of chemistry at age 29. He won the Nobel Prize in Chemistry in 1913 for his work on coordination compounds,
which he performed as a graduate student and first presented at age 26. Apparently, Werner was so obsessed with solving the
riddle of the structure of coordination compounds that his brain continued to work on the problem even while he was asleep. In
1891, when he was only 25, he woke up in the middle of the night and, in only a few hours, had laid the foundation for modern
coordination chemistry.

Table 3.1.1 : Werner’s Data on Complexes of Ammonia with P tC l 4

Number of Cl− Ions Precipitated


Complex Conductivity (ohm−1) Number of Ions per Formula Unit
by Ag+

PtCl4·6NH3 523 5 4

PtCl4·5NH3 404 4 3

PtCl4·4NH3 299 3 2

PtCl4·3NH3 97 2 1

PtCl4·2NH3 0 0 0

These data led Werner to postulate that metal ions have two different kinds of valence: (1) a primary valence (oxidation state) that
corresponds to the positive charge on the metal ion and (2) a secondary valence (coordination number) that is the total number of
ligand-metal bonds bound to the metal ion. If Pt had a primary valence of 4 and a secondary valence of 6, Werner could explain
the properties of the PtCl ⋅ NH adducts by the following reactions, where the metal complex is enclosed in square brackets:
4 3

4+ −
[Pt(NH3 )6 ]C l4 → [Pt(NH3 )6 ] (aq) + 4C l (aq)

3+ −
[Pt(NH3 )5 Cl]C l3 → [Pt(NH3 )5 Cl ] (aq) + 3C l (aq)

2+ −
[Pt(NH3 )4 C l2 ]C l2 → [Pt(NH3 )4 C l2 ] (aq) + 2C l (aq)

+ −
[Pt(NH3 )3 C l3 ]Cl → [Pt(NH3 )3 C l3 ] (aq) + C l (aq)

0
[Pt(NH3 )2 C l4 ] → [Pt(NH3 )2 C l4 ] (aq)

Further work showed that the two missing members of the series—[Pt(NH3)Cl5]− and [PtCl6]2−—could be prepared as their mono-
and dipotassium salts, respectively. Similar studies established coordination numbers of 6 for Co3+ and Cr3+ and 4 for Pt2+ and
Pd2+.
Werner’s studies on the analogous Co3+ complexes also allowed him to propose a structural model for metal complexes with a
coordination number of 6. Thus he found that [Co(NH3)6]Cl3 (yellow) and [Co(NH3)5Cl]Cl2 (purple) were 1:3 and 1:2 electrolytes.
Unexpectedly, however, two different [Co(NH3)4Cl2]Cl compounds were known: one was red, and the other was green (Figure
3.1.1a). Because both compounds had the same chemical composition and the same number of groups of the same kind attached to

the same metal, there had to be something different about the arrangement of the ligands around the metal ion. Werner’s key insight
was that the six ligands in [Co(NH3)4Cl2]Cl had to be arranged at the vertices of an octahedron because that was the only structure
consistent with the existence of two, and only two, arrangements of ligands (Figure 3.1.1b. His conclusion was corroborated by the
existence of only two different forms of the next compound in the series: Co(NH3)3Cl3.

3.1.2 https://chem.libretexts.org/@go/page/48465
Figure 3.1.1 : Complexes with Different Arrangements of the Same Ligands Have Different Properties. The [Co(NH3)4Cl2]+ ion
can have two different arrangements of the ligands, which results in different colors: if the two Cl− ligands are next to each other,
the complex is red (a), but if they are opposite each other, the complex is green (b).

 Example 3.1.1

In Werner’s time, many complexes of the general formula MA4B2 were known, but no more than two different compounds
with the same composition had been prepared for any metal. To confirm Werner’s reasoning, calculate the maximum number
of different structures that are possible for six-coordinate MA4B2 complexes with each of the three most symmetrical possible
structures: a hexagon, a trigonal prism, and an octahedron. What does the fact that no more than two forms of any MA4B2
complex were known tell you about the three-dimensional structures of these complexes?
Given: three possible structures and the number of different forms known for MA4B2 complexes
Asked for: number of different arrangements of ligands for MA4B2 complex for each structure
Strategy:
Sketch each structure, place a B ligand at one vertex, and see how many different positions are available for the second B
ligand.
Solution
The three regular six-coordinate structures are shown here, with each coordination position numbered so that we can keep track
of the different arrangements of ligands. For each structure, all vertices are equivalent. We begin with a symmetrical MA6
complex and simply replace two of the A ligands in each structure to give an MA4B2 complex:

For the hexagon, we place the first B ligand at position 1. There are now three possible places for the second B ligand: at
position 2 (or 6), position 3 (or 5), or position 4. These are the only possible arrangements. The (1, 2) and (1, 6) arrangements
are chemically identical because the two B ligands are adjacent to each other. The (1, 3) and (1, 5) arrangements are also
identical because in both cases the two B ligands are separated by an A ligand.
Turning to the trigonal prism, we place the first B ligand at position 1. Again, there are three possible choices for the second B
ligand: at position 2 or 3 on the same triangular face, position 4 (on the other triangular face but adjacent to 1), or position 5 or
6 (on the other triangular face but not adjacent to 1). The (1, 2) and (1, 3) arrangements are chemically identical, as are the (1,
5) and (1, 6) arrangements.
In the octahedron, however, if we place the first B ligand at position 1, then we have only two choices for the second B ligand:
at position 2 (or 3 or 4 or 5) or position 6. In the latter, the two B ligands are at opposite vertices of the octahedron, with the
metal lying directly between them. Although there are four possible arrangements for the former, they are chemically identical
because in all cases the two B ligands are adjacent to each other.

3.1.3 https://chem.libretexts.org/@go/page/48465
The number of possible MA4B2 arrangements for the three geometries is thus: hexagon, 3; trigonal prism, 3; and octahedron, 2.
The fact that only two different forms were known for all MA4B2 complexes that had been prepared suggested that the correct
structure was the octahedron but did not prove it. For some reason one of the three arrangements possible for the other two
structures could have been less stable or harder to prepare and had simply not yet been synthesized. When combined with
analogous results for other types of complexes (e.g., MA3B3), however, the data were best explained by an octahedral structure
for six-coordinate metal complexes.

 Exercise 3.1.1
Determine the maximum number of structures that are possible for a four-coordinate MA2B2 complex with either a square
planar or a tetrahedral symmetrical structure.

Answer
square planar, 2; tetrahedral, 1

Structures of Metal Complexes


The coordination numbers of metal ions in metal complexes can range from 2 to at least 9. In general, the differences in energy
between different arrangements of ligands are greatest for complexes with low coordination numbers and decrease as the
coordination number increases. Usually only one or two structures are possible for complexes with low coordination numbers,
whereas several different energetically equivalent structures are possible for complexes with high coordination numbers (n > 6).
The following presents the most commonly encountered structures for coordination numbers 2–9. Many of these structures should
be familiar to you from our discussion of the valence-shell electron-pair repulsion (VSEPR) model because they correspond to the
lowest-energy arrangements of n electron pairs around a central atom.

Compounds with low coordination numbers exhibit the greatest differences in energy
between different arrangements of ligands.
Coordination Number 2
Although it is rare for most metals, this coordination number is surprisingly common for d10 metal ions, especially Cu+, Ag+, Au+,
and Hg2+. An example is the [Au(CN)2]− ion, which is used to extract gold from its ores. As expected based on VSEPR
considerations, these complexes have the linear L–M–L structure shown here.

Coordination Number 3
Although it is also rare, this coordination number is encountered with d10 metal ions such as Cu+ and Hg2+. Among the few known
examples is the HgI3− ion. Three-coordinate complexes almost always have the trigonal planar structure expected from the VSEPR
model.

Coordination Number 4
Two common structures are observed for four-coordinate metal complexes: tetrahedral and square planar. The tetrahedral structure
is observed for all four-coordinate complexes of nontransition metals, such as [BeF4]2−, and d10 ions, such as [ZnCl4]2−. It is also
found for four-coordinate complexes of the first-row transition metals, especially those with halide ligands (e.g., [FeCl4]− and
[FeCl4]2−). In contrast, square planar structures are routinely observed for four-coordinate complexes of second- and third-row

3.1.4 https://chem.libretexts.org/@go/page/48465
transition metals with d8 electron configurations, such as Rh+ and Pd2+, and they are also encountered in some complexes of Ni2+
and Cu2+.

Coordination Number 5
This coordination number is less common than 4 and 6, but it is still found frequently in two different structures: trigonal
bipyramidal and square pyramidal. Because the energies of these structures are usually rather similar for most ligands, many five-
coordinate complexes have distorted structures that lie somewhere between the two extremes.

Coordination Number 6
This coordination number is by far the most common. The six ligands are almost always at the vertices of an octahedron or a
distorted octahedron. The only other six-coordinate structure is the trigonal prism, which is very uncommon in simple metal
complexes.

Coordination Number 7
This relatively uncommon coordination number is generally encountered for only large metals (such as the second- and third-row
transition metals, lanthanides, and actinides). At least three different structures are known, two of which are derived from an
octahedron or a trigonal prism by adding a ligand to one face of the polyhedron to give a “capped” octahedron or trigonal prism. By
far the most common, however, is the pentagonal bipyramid.

3.1.5 https://chem.libretexts.org/@go/page/48465
Coordination Number 8
This coordination number is relatively common for larger metal ions. The simplest structure is the cube, which is rare because it
does not minimize interligand repulsive interactions. Common structures are the square antiprism and the dodecahedron, both of
which can be generated from the cube.

Coordination Number 9
This coordination number is found in larger metal ions, and the most common structure is the tricapped trigonal prism, as in
[Nd(H2O)9]3+.

Key Takeaways
Coordination compounds are a major feature of the chemistry of over half the elements.
Coordination compounds have important roles as industrial catalysts in controlling reactivity, and they are essential in
biochemical processes.

Summary
Transition metals form metal complexes, polyatomic species in which a metal ion is bound to one or more ligands, which are
groups bound to a metal ion. Complex ions are electrically charged metal complexes, and a coordination compound contains one or
more metal complexes. Metal complexes with low coordination numbers generally have only one or two possible structures,
whereas those with coordination numbers greater than six can have several different structures. Coordination numbers of two and
three are common for d10 metal ions. Tetrahedral and square planar complexes have a coordination number of four; trigonal
bipyramidal and square pyramidal complexes have a coordination number of five; and octahedral complexes have a coordination
number of six. At least three structures are known for a coordination number of seven, which is generally found for only large
metal ions. Coordination numbers of eight and nine are also found for larger metal ions.

3.1.6 https://chem.libretexts.org/@go/page/48465
3.1: Werner’s Theory of Coordination Compounds is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.
24.1: Werner’s Theory of Coordination Compounds by Anonymous is licensed CC BY-NC-SA 4.0.

3.1.7 https://chem.libretexts.org/@go/page/48465
3.2: Ligands
A metal ion in solution does not exist in isolation, but in combination with ligands (such as solvent molecules or simple ions) or
chelating groups, giving rise to complex ions or coordination compounds. These complexes contain a central atom or ion, often a
transition metal, and a cluster of ions or neutral molecules surrounding it. Ligands are ions or neutral molecules that bond to a
central metal atom or ion. Ligands act as Lewis bases (electron pair donors), and the central atom acts as a Lewis acid (electron pair
acceptor). Ligands have at least one donor atom with an electron pair used to form covalent bonds with the central atom.
The term ligand come from the latin word ligare (which meaning to bind) was first used by Alfred Stock in 1916 in relation to
silicon chemistry. Ligands can be anions, cations, or neutral molecules. Ligands can be further characterized as monodentate,
bidentate, tridentate etc. where the concept of teeth (dent) is introduced, hence the idea of bite angle etc. A monodentate ligand has
only one donor atom used to bond to the central metal atom or ion.

Monodentate Ligands
The term "monodentate" can be translated as "one tooth," referring to the ligand binding to the center through only one atom. Some
examples of monodentate ligands are: chloride ions (referred to as chloro when it is a ligand), water (referred to as aqua when it is a
ligand), hydroxide ions (referred to as hydroxo when it is a ligand), and ammonia (referred to as ammine when it is a ligand).

Figure 3.2.1 : Central atom with six monodentate ligands attached. (Image courtesy of Wikimedia Commons.)

Bidentate Ligands
Bidentate ligands have two donor atoms which allow them to bind to a central metal atom or ion at two points. Common examples
of bidentate ligands are ethylenediamine (en), and the oxalate ion (ox). Shown below is a diagram of ethylenediamine: the nitrogen
(blue) atoms on the edges each have two free electrons that can be used to bond to a central metal atom or ion.

Figure 3.2.2 : Ethylenediamine (en) is an example of a bidentate ligand. (Image courtesy of Wikimedia Commons.).
Ethylenediamine (en) is a bidentate ligand that forms a five-membered ring in coordinating to a metal ion M

Polydentate Ligands
range in the number of atoms used to bond to a central metal atom or ion. EDTA, a
Polydentate ligands
hexadentate ligand, is an example of a polydentate ligand that has six donor atoms with electron pairs
that can be used to bond to a central metal atom or ion.

3.2.1 https://chem.libretexts.org/@go/page/48466
Figure 3.2.3 : Ethylenediaminetetraaceticacid acid (EDTA) is a hexadentate ligand and can bind a Metal via multiple "teeth" (left)
Unbound EDTA ion and (right) EDTA bound to a generic transition metal. (Images courtesy of Wikimedia Commons.)
Unlike polydentate ligands, ambidentate ligands can attach to the central atom in two places. A good example of this is thiocyanate,
SC N , which can attach at either the sulfur atom or the nitrogen atom.

Example 3.2.1

Draw metal complexes using the ligands below and metal ions of your choice.

Chelation
Chelation is a process in which a polydentate ligand bonds to a metal ion, forming a ring. The complex produced by this process is
called a chelate, and the polydentate ligand is referred to as a chelating agent. The term chelate was first applied in 1920 by Sir
Gilbert T. Morgan and H.D.K. Drew, who stated: "The adjective chelate, derived from the great claw or chela (chely- Greek) of the
lobster or other crustaceans, is suggested for the caliperlike groups which function as two associating units and fasten to the central
atom so as to produce heterocyclic rings." As the name implies, chelating ligands have high affinity for metal ions relative to
ligands with only one binding group (which are called monodentate = "single tooth") ligands. Both Ethylenediamine (Figure 3.2.2)
and Ethylenediaminetetraaceticacid acid (Figure 3.2.3) are examples of chelating agents, but many others are commonly found in
the inorganic laboratory.

3.2.2 https://chem.libretexts.org/@go/page/48466
The Chelate Effect
The chelate effect is the enhanced affinity of chelating ligands for a metal ion compared to the affinity of a collection of similar
nonchelating (monodentate) ligands for the same metal.

The macrocyclic effect follows the same principle as the chelate effect, but the effect is further enhanced by the cyclic
conformation of the ligand. Macrocyclic ligands are not only multi-dentate, but because they are covalently constrained to their
cyclic form, they allow less conformational freedom. The ligand is said to be "pre-organized" for binding, and there is little entropy
penalty for wrapping it around the metal ion. For example heme b is a tetradentate cyclic ligand that strongly complexes transition
metal ions, including Fe+2 in hemogloben (Figure 3.2.4).

Figure 3.2.4 : Heme b macrocyclic ligand that binds iron in hemoglobin


Some other common chelating and cyclic ligands are shown below:
Acetylacetonate (acac-, top) is an anionic bidentate ligand that coordinates metal ions through two oxygen atoms. Acac- is a
hard base so it prefers hard acid cations. With divalent metal ions, acac-forms neutral, volatile complexes such as Cu(acac)2 and
Mo(acac)2 that are useful for chemical vapor deposition (CVD) of metal thin films.

2,2'-Bipyridine 2,2'-Bipyridine (Figure 3.2.5: left) and related bidentate ligands such as 1,10-phenanthroline form propeller-
shaped complexes with metals such as Ru2+. The [Ru(bpy)3]2+ complex is photoluminescent and can also undergo photoredox
reactions, making it an interesting compound for both photocatalysis and artificial photosynthesis.
Crown ethers such as 18-crown-6 2,2'-Bipyridine (Figure 3.2.5: center) are cyclic hard bases that can complex alkali metal
cations. Crowns can selectively bind Li+, Na+, or K+ depending on the number of ethylene oxide units in the ring.
The chelating properties of crown ethers are mimetic of the natural antibiotic valinomycin ( 2,2'-Bipyridine (Figure 3.2.5:
right), which selectively transports K+ ions across bacterial cell membranes, killing the bacterium by dissipating its membrane
potential. Like crown ethers, valinomycin is a cyclic hard base.

Figure 3.2.5 : (let) 2,2'-Bipyridine, (center) 18-crown-6 2,2'-Bipyridine (right) valinomycin

References
1. Petrucci, Harwood, Herring, Madura. General Chemistry Principles & Modern Applications. Prentice Hall. New Jersey, 2007
2. Cox, Tony. (2004). Instant notes in inorganic chemistry. Oxford, UK: Taylor & Francis.
3. Libraries, Association, Robert Williams, and J. Silva. Bringing chemistry to life. Oxford University Press, USA, 1999. Print.
4. Moeller, Therald, Douville, Judith, & Libraries, Association. (1988). Inorganic Chemistry: A Modern Introduction. Amer
Library Assn.
5. Bowker, R., Warmus, Mieczysław, Muzzy, Adrienne, LOCALIZADO, AUTOR, Hopkinson, Barbara, Saur, K, Izod, Irene,
Hopkinson, Barbara, Saur, K, Books, K, & Company, K. (1994). Inorganic Chemistry Concepts. K G Saur Verlag Gmbh & Co.
6. Porterfield, William. (1984). Inorganic chemistry. Addison Wesley Publishing Company.

3.2.3 https://chem.libretexts.org/@go/page/48466
Contributors and Attributions
Harjeet Bassi (UCD)
Adapted from the Wikibook constructed by Chemistry 310 students at Penn State University.

3.2: Ligands is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

3.2.4 https://chem.libretexts.org/@go/page/48466
3.3: Nomenclature
 Learning Objectives
To learn the basis for complex ion and compound nomenclature

Coordination complexes have their own classes of isomers, different magnetic properties and colors, and various applications
(photography, cancer treatment, etc), so it makes sense that they would have a naming system as well. Consisting of a metal and
ligands, their formulas follow the pattern [Metal ligands]±Charge, while names are written Prefix Ligands Metal (Oxidation State).

Introduction
According to the Lewis base theory, ligands are Lewis bases since they can donate electrons to the central metal atom. The metals,
in turn, are Lewis acids since they accept electrons. Coordination complexes consist of a ligand and a metal center cation. The
overall charge can be positive, negative, or neutral. Coordination compounds are complex or contain complex ions, for example:
Complex Cation: [Co(NH ) ] 3 6
3 +


Complex Anion: [CoCl (NH ) ] 4 3 2

Neutral Complex: [CoCl (NH ) ] 3 3 3

Coordination Compound: K [Fe(CN) 4 6


]

A ligand can be an anion or a neutral molecule that donates an electron pair to the complex (NH3, H2O, Cl-). The number of
ligands that attach to a metal depends on whether the ligand is monodentate or polydentate. To begin naming coordination
complexes, here are some things to keep in mind.
1. Ligands are named first in alphabetical order.
2. The name of the metal comes next.
3. The oxidation state of the metal follows, noted by a Roman numeral in parentheses (II, IV).

Rule 1: Anionic Ligands


Ligands that act as anions which end in "-ide" are replaced with an ending "-o" (e.g., Chloride → Chloro). Anions ending with "-
ite" and "-ate" are replaced with endings "-ito" and "-ato" respectively (e.g., Nitrite → Nitrito, Nitrate → Nitrato).
Table 3.3.1 : Anionic Monodentate Ligands
Molecular Formula Ligand Name Molecular Formula Ligand Name

F- Fluoro OH- Hydroxo


-
Cl Chloro SO42- Sulfato
-
Br Bromo S2O32- Thiosulfato

I- Iodo NO2- Nitrito-N-; Nitro


2- -
O Oxo ONO Nitrito-O-; Nitrito
- -
CN Cyano SCN Thiocyanato-S-; Thiocyanato

NC- Isocyano NCS- Thiocyanato-N-; Isothiocyanato

Rule 2: Neutral Ligands


Most neutral molecules that are ligands carry their normal name. The few exceptions are the first four on the chart: ammine, aqua,
carbonyl, and nitrosyl.
Table 3.3.2 : Select Neutral Monodentate Ligands. Note: Ammine is spelled with two m's when referring to a ligand. Amines are a class of
organic nitrogen-containing compounds.
Molecular Formula of Ligand Ligand Name

NH3 Ammine

H2O Aqua

CO Carbonyl

3.3.1 https://chem.libretexts.org/@go/page/48467
Molecular Formula of Ligand Ligand Name

NO Nitrosyl

CH3NH2 Methylamine

C5H5N Pyridine

Polydentate ligands follow the same rules for anions and neutral molecules.
Table 3.3.3 : Select Polydentate ligands
Short name Extended name

en Ethylenediamine

ox2- Oxalato

EDTA4- Ethylenediaminetetraacetato

Rule 3: Ligand Multiplicity


The number of ligands present in the complex is indicated with the prefixes di, tri, etc. The exceptions are polydentates that have a
prefix already in their name (en and EDTA4- are the most common). When indicating how many of these are present in a
coordination complex, put the ligand's name in parentheses and use bis (for two ligands), tris (for three ligands), and tetrakis (for
four ligands).
Table 3.3.4 : Prefixes for indicating number of ligands in a complex.
Number of Ligands Monodentate Ligands Polydentate Ligands

1 mono -

2 di bis

3 tri tris

4 tetra tetrakis

5 penta pentakis

6 hexa hexakis

Prefixes always go before the ligand name; they are not taken into account when putting ligands in alphabetical order. Note that
"mono" often is not used. For example, [FeCl(CO) (NH ) ] would be called triamminedicarbonylchloroiron(III) ion.
2 3 3
2 +

Remember that ligands are always named first, before the metal is.

 Example 3.3.1

What is the name of this complex ion: [CrCl 2


(H O) ]
2 4
+
?

Solution
Let's start by identifying the ligands. The ligands here are Cl and H2O. Therefore, we will use the monodentate ligand names of
"chloro" and "aqua". Alphabetically, aqua comes before chloro, so this will be their order in the complex's name. There are 4
aqua's and 2 chloro's, so we will add the number prefixes before the names. Since both are monodentate ligands, we will say
"tetra[aqua]di[chloro]".
Now that the ligands are named, we will name the metal itself. The metal is Cr, which is chromium. Therefore, this
coordination complex is called tetraaquadichlorochromium(III) ion. See the next section for an explanation of the (III).

3.3.2 https://chem.libretexts.org/@go/page/48467
 Exercise 3.3.1

What is the name of this complex ion: [CoCl 2


(en) ]
2
+
?

Answer
We take the same approach. There are two chloro and ethylenediamine ligands. The metal is Co, cobalt. We follow the
same steps, except that en is a polydentate ligand with a prefix in its name (ethylenediamine), so "bis" is used instead of
"di", and parentheses are added. Therefore, this coordination complex is called dichlorobis(ethylenediamine)cobalt(III) ion.

Rule 4: The Metals


When naming the metal center, you must know the formal metal name and the oxidation state. To show the oxidation state, we use
Roman numerals inside parenthesis. For example, in the problems above, chromium and cobalt have the oxidation state of +3, so
that is why they have (III) after them. Copper, with an oxidation state of +2, is denoted as copper(II). If the overall coordination
complex is an anion, the ending "-ate" is attached to the metal center. Some metals also change to their Latin names in this
situation. Copper +2 will change into cuprate(II). The following change to their Latin names when part of an anion complex:
Table 3.3.5 : Latin terms for Select Metal Ion
Transition Metal Latin

Iron Ferrate

Copper Cuprate

Tin Stannate

Silver Argentate

Lead Plumbate

Gold Aurate

The rest of the metals simply have -ate added to the end (cobaltate, nickelate, zincate, osmate, cadmate, platinate, mercurate, etc.
Note that the -ate tends to replace -um or -ium, if present).
Finally, when a complex has an overall charge, "ion" is written after it. This is not necessary if it is neutral or part of a coordination
compound (Example 3.3.3). Here are some examples with determining oxidation states, naming a metal in an anion complex, and
naming coordination compounds.

 Example 3.3.2

What is the name of [Cr(OH)4]- ?

Solution
Immediately we know that this complex is an anion. There is only one monodentate ligand, hydroxide. There are four of them,
so we will use the name "tetrahydroxo". The metal is chromium, but since the complex is an anion, we will have to use the "-
ate" ending, yielding "chromate". The oxidation state of the metal is 3 (x+(-1)4=-1). Write this with Roman numerals and
parentheses (III) and place it after the metal to get tetrahydroxochromate(III) ion.

 Exercise 3.3.2

What is the name of [CuCl 4


2 −
] ?

Answer
tetrachlorocuprate(II) ion

3.3.3 https://chem.libretexts.org/@go/page/48467
A last little side note: when naming a coordination compound, it is important that you name the cation first, then the anion. You
base this on the charge of the ligand. Think of NaCl. Na, the positive cation, comes first and Cl, the negative anion, follows.

 Example 3.3.3

What is the name of [Pt(NH 3


) ][Pt(Cl) ]
4 4
?

Solution
NH3 is neutral, making the first complex positively charged overall. Cl has a -1 charge, making the second complex the anion.
Therefore, you will write the complex with NH3 first, followed by the one with Cl (the same order as the formula). This
coordination compound is called tetraammineplatinum(II) tetrachloroplatinate(II).

 Exercise 3.3.3: The Nitro/Nitrito Ambidentate Ligand

What is the name of [CoCl(NO 2


)(NH ) ]
3 4
+
?

Answer
This coordination complex is called tetraamminechloronitrito-N-cobalt(III). N comes before the O in the symbol for the
nitrite ligand, so it is called nitrito-N. If an O came first, as in [CoCl(ONO)(NH3)4]+, the ligand would be called nitrito-O,
yielding the name tetraamminechloronitrito-O-cobalt(III).
Nitro (for NO2) and nitrito (for ONO) can also be used to describe the nitrite ligand, yielding the names
tetraamminechloronitrocobalt(III) and tetraamminechloronitritocobalt(III).

Writing Formulas of Coordination Complexes


While chemistry typically follow the nomenclature rules for naming complexes and compounds, there is disagreement with the
rules for constructing formulas of inorganic complex. The order of ligand names in their formula has been ambiguous with
different conventions being used (charged vs neutral, number of each ligand, etc.). In 2005, IUPAC adopted the recommendation
that all ligand names in formulas be listed alphabetically (in the same way as in the naming convention) irrespective of the charge
or number of each ligand type.
However, this rule is not adhered to in many chemistry laboratories. For practice, the order of the ligands in chemical formulas
does not matter as long as you write the transition metal first, which is the stance taken here.

 Examples 3.3.4

Write the chemical formulas for:


a. Amminetetraaquachromium(II) ion
b. Amminesulfatochromium(II)

Solution
+2 +2
a. Amminetetraaquachromium(II) ion could be written as [Cr(H O) (NH )] or [Cr(NH
2 4 3 3
)(H O) ]
2 4
.
b. Amminesulfatochromium (II) could be written as [Cr(SO )(NH )] or [Cr(NH )(SO )].
4 3 3 4

 Exercise 3.3.4

Write the chemical formulas for


a. Amminetetraaquachromium (II) sulfate
b. Potassium hexacyanoferrate (III)

Answer

3.3.4 https://chem.libretexts.org/@go/page/48467
a. Amminetetraaquachromium (II) sulfate can be written as [Cr(H 2
O) (NH )] SO
4 3 4
. Although [Cr(NH 3
)(H O) ] SO
2 4 4
is
also acceptable.
b. Potassium hexacyanoferrate (III) is be written as K [Fe(CN) ]
3 6

References
1. Petrucci, Ralph H. General Chemistry Principles and Modern Applications. 9th ed. Upper Saddle River: Pearson Prentice Hall,
2002.

Contributors and Attributions


Justin Hosung Lee (UCD), Sophia Muller (UCD)

3.3: Nomenclature is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
23.3: Nomenclature of Coordination Chemistry is licensed CC BY-NC-SA 3.0.

3.3.5 https://chem.libretexts.org/@go/page/48467
3.4: Isomerism
Learning Objectives
Explain the differences of Structural and Geometric isomerization in a coordination complexes or complex ions.
Define Ionization, Linkage, and Coordination Isomerization (structural isomer classes)
Define cis/tran and mer/fac isomerization (geometric isomer classes)

The existence of coordination compounds with the same formula but different arrangements of the ligands was crucial in the
development of coordination chemistry. Two or more compounds with the same formula but different arrangements of the atoms
are called isomers. Because isomers usually have different physical and chemical properties, it is important to know which isomer
we are dealing with if more than one isomer is possible. As we will see, coordination compounds exhibit the same types of isomers
as organic compounds, as well as several kinds of isomers that are unique. Isomers are compounds with the same molecular
formula but different structural formulas and do not necessarily share similar properties. There are many different classes of
isomers, like stereoisomers, enantiomers, and geometrical isomers. There are two main forms of isomerism: structural isomerism
and stereoisomerism (spatial isomerism).

Class I: Structural Isomers


Isomers that contain the same number of atoms of each kind but differ in which atoms are bonded to one another are called
structural isomers, which differ in structure or bond type. For inorganic complexes, there are three types of structural isomers:
ionization, coordination and linkage. Structural isomers, as their name implies, differ in their structure or bonding, which are
separate from stereoisomers that differ in the spatial arrangement of the ligands are attached, but still have the bonding properties.
The different chemical formulas in structural isomers are caused either by a difference in what ligands are bonded to the central
atoms or how the individual ligands are bonded to the central atoms. When determining a structural isomer, you look at (1) the
ligands that are bonded to the central metal and (2) which atom of the ligands attach to the central metal.

Ionization Isomerism
Ionization isomers occur when a ligand that is bound to the metal center exchanges places with an anion or neutral molecule that
was originally outside the coordination complex. The geometry of the central metal ion and the identity of other ligands are
identical. For example, an octahedral isomer will have five ligands that are identical, but the sixth will differ. The non-matching
ligand in one compound will be outside of the coordination sphere of the other compound. Because the anion or molecule outside
the coordination sphere is different, the chemical properties of these isomers is different. A hydrate isomer is a specific kind of
ionization isomer where a water molecule is one of the molecules that exchanges places.

Figure 3.4.1 : Ionization isomerism. The two isomers differ only in which ligands are bound to the center metal and which are
counter ions in the second coordination sphere. (left) The chloride ion is bound to the cobalt as a chloro-ligand with the bromide
ion as the counter ion. (right) In the other ionization isomer the bromide is acting as a bromo-ligand to the cobalt with the chloride
acting as the counter ion. These two isomers are called pentaamminechlorocobalt(II) bromide (left) and
pentaamminebromocobalt(II) chloride (right).
The difference between the ionization isomers can be view within the context of the ions generated when each are dissolved in
solution. For example, when pentaamminebromocobalt(II) chloride is dissolved in water, Cl ions are generated:

+ −
CoBr(NH ) Cl(s) → CoBr(NH ) (aq) + Cl (aq) (3.4.1)
3 5 3 5

whereas when pentaamminechlorocobalt(II) bromide is dissolved, Br ions are generated:


+ −
CoCl (NH ) Br(s) → CoCl (NH ) (aq) + Br (aq) (3.4.2)
3 5 3 5

Notice that both anions are necessary to balance the charge of the complex, and that they differ in that one ion is directly attached
to the central metal, but the other is not.

3.4.1 https://chem.libretexts.org/@go/page/48468
Solvate and Hydrate Isomerism: A special Type of Ionization Isomers
A very similar type of isomerism results from replacement of a coordinated group by a solvent molecule (Solvate Isomerism),
which in the case of water is called Hydrate Isomerism. The best known example of this occurs for chromium chloride (
CrCl ⋅ 6 H O ) which may contain 4, 5, or 6 coordinated water molecules (assuming a coordination number of 6). The dot
3 2

here is used essentially as an expression of ignorance to indicate that, though the parts of the molecule separated by the dot are
bonded to one another in some fashion, the exact structural details of that interaction are not fully expressed in the resulting
formula. Alfred Werner’s coordination theory indicates that several of the water molecules are actually bonded directly (via
coordinate covalent bonds) to the central chromium ion. In fact, there are several possible compounds that use the brackets to
signify bonding in the complex and the dots to signify "water molecules that are not bound to the central metal, but are part of
the lattice:
[ CrCl (H O) ]Cl ⋅ 2 H O
2 2 4
: bright-green colored
2

[CrCl (H O) ] Cl
2 5 2 2
: grey-green colored
⋅H O

[Cr(H O) ] Cl
2 6 3
: violet colored
These isomers have very different chemical properties and on reaction with AgNO to test for Cl ions, would find 1, 2, and 3
3

Cl

ions in solution, respectively.
Upon crystallization from water, many compounds incorporate water molecules in their crystalline frameworks. These "waters
of crystallization" refers to water that is found in the crystalline framework of a metal complex or a salt, which is not directly
bonded to the metal cation. In the first two hydrate isomers, there are water molecules that are artifacts of the crystallization
and occur inside crystals. These waters of crystallization contribute to the total weight of water in a substance and are mostly
present in a definite (stoichiometric) ratio.

Figure 3.4.2 : Hydrate iosomerisim. The [Cr(H O) ]Cl


2 6 3
hydrate isomer (left) is violet colored and the
[CrCl( H O) ]Cl
2 5 2
⋅H O
2
hydrate isomer is green-grey colored.
What are "Waters of Crystallization"?
A compound with associated water of crystallization is known as a hydrate. The structure of hydrates can be quite elaborate,
because of the existence of hydrogen bonds that define polymeric structures. For example, consider the aquo complex
NiCl ⋅ 6 H O that consists of separated trans-[NiCl2(H2O)4] molecules linked more weakly to adjacent water molecules.
2 2

Only four of the six water molecules in the formula are bound to the nickel (II) cation, and the remaining two are waters of
crystallization as the crystal structure resolves (Figure 3.4.3).

3.4.2 https://chem.libretexts.org/@go/page/48468
Figure 3.4.3 : Water of hydration. (left) Image of NiCl ⋅ 6 H O salt (Public Domain; Benjah-bmm27 via Wikipedia). (right)
2 2

Crystal Structure of NiCl ⋅ 6 H O with chlorine atoms (green), water molecules (red), and Ni metals (blue) indicated. Note
2 2

that only four of the waters are bound as ligands to the nickel ions and two are outside of the coordination sphere. (CC BY-SA
4.0; Smokefoot).
Water is particularly common solvent to be found in crystals because it is small and polar. But all solvents can be found in
some host crystals. Water is noteworthy because it is reactive, whereas other solvents such as benzene are considered to be
chemically innocuous.

Coordination Isomerism
Coordination isomerism occurs in compounds containing complex anionic and complex cationic parts and can be viewed as
an interchange of some ligands from the cation to the anion. Hence, there are two complex compounds bound together, one with a
negative charge and the other with a positive charge. In coordination isomers, the anion and cation complexes of a coordination
compound exchange one or more ligands. For example, the [Zn(NH ) ][Cu(Cl )] and [Cu(NH ) ][Zn(Cl )] compounds are
3 4 4 3 4 4

coordination isomers (Figure 3.4.4).

Figure 3.4.4 : Coordination Isomerism. (CC BY-NC; Ümit Kaya)

Example 3.4.2: Coordination Isomers

What is the coordination isomer for the [Co(NH 3


) ][Cr(CN) ]
6 6
compound?
Solution
Coordination isomerism involves switching the metals between the cation and anion spheres.

Hence the [Cr(NH 3


) ][Co (CN) ]
6 6
compound is a coordination isomer of [Co(NH 3
) ][Cr(CN) ]
6 6
.

Linkage Isomerism
Linkage isomerism occurs with ambidentate ligands that are capable of coordinating in more than one way. The best known cases
involve the monodentate ligands: SCN /NCS and NO /ONO . The only difference is what atoms the molecular ligands
− − −

2

3.4.3 https://chem.libretexts.org/@go/page/48468
bind to the central ion. The ligand(s) must have more than one donor atom, but bind to ion in only one place. For example, the (
NO ) ion is a ligand that can bind to the central atom through the nitrogen or the oxygen atom, but cannot bind to the central atom

with both oxygen and nitrogen at once, in which case it would be called a polydentate ligand rather than an ambidentate ligand.

Figure 3.4.5 : Linkage Isomerism in the NO ligand. This occurs when a particular ligand is capable of coordinating to a metal in
2

two different and distinct ways. from Angel C. de Dios.


As with all structural isomers, the formula of the complex is unchanged for each isomer, but the properties may differ. The names
used to specify the changed ligands are changed as well. For example, the NO ion is called nitro when it binds with the N atom

2

and is called nitrito when it binds with the O atom.

Example 3.4.3: Linkage Isomerization

The cationic cobalt complex [Co(NH3)5(NO2)]Cl2 exists in two separable linkage isomers of the complex ion: (NH3)5(NO2)]2+.

Figure 3.4.6 : Linkage Isomerization. (left) The nitro isomer (Co-NO2) and (right) the nitrito isomer (Co-ONO)
When donation is from nitrogen to a metal center, the complex is known as a nitro- complex and when donation is from one
oxygen to a metal center, the complex is known as a nitrito- complex. An alternative formula structure to emphasize the
different coordinate covalent bond for the two isomers
[Co(ONO)(NH ) ]Cl
3 5
: the nitrito isomer -O attached
[Co(NO )(NH ) ]Cl
2 3 5
: the nitro isomer - N attached.

Another example of an ambidentate ligand is thiocyanate, SCN , which can attach to the transition metal at either the sulfur atom

or the nitrogen atom. Such compounds give rise to linkage isomerism. Other ligands that give rise to linkage isomers include
selenocyanate, SeCN – isoselenocyanate, NCSe and sulfite, SO .
− − 2 −
3

Exercise 3.4.3
3 – 3 –
Are [FeCl 5
(NO )]
2
and [FeCl5
(ONO)] complex ions linkage isomers of each other?

Answer
Here, the difference is in how the ligand bonds to the metal. In the first isomer, the ligand bonds to the metal through an
electron pair on the nitrogen. In the second isomer, the ligand bonds to the metal through an electron pair on one of the
oxygen atoms. It's easier to see it:

3.4.4 https://chem.libretexts.org/@go/page/48468
Class 2: Geometric Isomers
The existence of coordination compounds with the same formula but different arrangements of the ligands was crucial in the
development of coordination chemistry. Two or more compounds with the same formula but different arrangements of the atoms
are called isomers. Because isomers usually have different physical and chemical properties, it is important to know which isomer
we are dealing with if more than one isomer is possible. Recall that in many cases more than one structure is possible for organic
compounds with the same molecular formula; examples discussed previously include n-butane versus isobutane and cis-2-butene
versus trans-2-butene. As we will see, coordination compounds exhibit the same types of isomers as organic compounds, as well as
several kinds of isomers that are unique.

Planar Isomers
Metal complexes that differ only in which ligands are adjacent to one another (cis) or directly across from one another (trans) in
the coordination sphere of the metal are called geometrical isomers. They are most important for square planar and octahedral
complexes.
Because all vertices of a square are equivalent, it does not matter which vertex is occupied by the ligand B in a square planar
MA3B complex; hence only a single geometrical isomer is possible in this case (and in the analogous MAB3 case). All four
structures shown here are chemically identical because they can be superimposed simply by rotating the complex in space:

For an MA2B2 complex, there are two possible isomers: either the A ligands can be adjacent to one another (cis), in which case the
B ligands must also be cis, or the A ligands can be across from one another (trans), in which case the B ligands must also be trans.
Even though it is possible to draw the cis isomer in four different ways and the trans isomer in two different ways, all members of
each set are chemically equivalent:

Because there is no way to convert the cis structure to the trans by rotating or flipping the molecule in space, they are
fundamentally different arrangements of atoms in space. Probably the best-known examples of cis and trans isomers of an MA2B2
square planar complex are cis-Pt(NH3)2Cl2, also known as cisplatin, and trans-Pt(NH3)2Cl2, which is actually toxic rather than
therapeutic.

The anticancer drug cisplatin and its inactive trans isomer. Cisplatin is especially effective against tumors of the reproductive
organs, which primarily affect individuals in their 20s and were notoriously difficult to cure. For example, after being diagnosed

3.4.5 https://chem.libretexts.org/@go/page/48468
with metastasized testicular cancer in 1991 and given only a 50% chance of survival, Lance Armstrong was cured by treatment
with cisplatin.
Square planar complexes that contain symmetrical bidentate ligands, such as [Pt(en)2]2+, have only one possible structure, in which
curved lines linking the two N atoms indicate the ethylenediamine ligands:

Exercise 3.4.4

Draw the cis and trans isomers of the following compounds:


a. (NH3)2IrCl(CO)
b. (H3P)2PtHBr
c. (AsH3)2PtH(CO)

Exercise 3.4.5

Only one isomer of (tmeda)PtCl2 is possible [tmeda = (CH3)2NCH2CH2N(CH3)2; both nitrogens connect to the platinum].
Draw this isomer and explain why the other isomer is not possible.

Octahedral Isomers
Octahedral complexes also exhibit cis and trans isomers. Like square planar complexes, only one structure is possible for
octahedral complexes in which only one ligand is different from the other five (MA5B). Even though we usually draw an
octahedron in a way that suggests that the four “in-plane” ligands are different from the two “axial” ligands, in fact all six vertices
of an octahedron are equivalent. Consequently, no matter how we draw an MA5B structure, it can be superimposed on any other
representation simply by rotating the molecule in space. Two of the many possible orientations of an MA5B structure are as
follows:

If two ligands in an octahedral complex are different from the other four, giving an MA4B2 complex, two isomers are possible. The
two B ligands can be cis or trans. Cis- and trans-[Co(NH3)4Cl2]Cl are examples of this type of system:

Replacing another A ligand by B gives an MA3B3 complex for which there are also two possible isomers. In one, the three ligands
of each kind occupy opposite triangular faces of the octahedron; this is called the fac isomer (for facial). In the other, the three

3.4.6 https://chem.libretexts.org/@go/page/48468
ligands of each kind lie on what would be the meridian if the complex were viewed as a sphere; this is called the mer isomer (for
meridional):

Example 3.4.4

Draw all the possible geometrical isomers for the complex [Co(H2O)2(ox)BrCl]−, where ox is −O2CCO2−, which stands for
oxalate.
Solution
Given: formula of complex
Asked for: structures of geometrical isomers
This complex contains one bidentate ligand (oxalate), which can occupy only adjacent (cis) positions, and four monodentate
ligands, two of which are identical (H2O). The easiest way to attack the problem is to go through the various combinations of
ligands systematically to determine which ligands can be trans. Thus either the water ligands can be trans to one another or the
two halide ligands can be trans to one another, giving the two geometrical isomers shown here:

In addition, two structures are possible in which one of the halides is trans to a water ligand. In the first, the chloride ligand is
in the same plane as the oxalate ligand and trans to one of the oxalate oxygens. Exchanging the chloride and bromide ligands
gives the other, in which the bromide ligand is in the same plane as the oxalate ligand and trans to one of the oxalate oxygens:

This complex can therefore exist as four different geometrical isomers.

Exercise 3.4.6

Draw all the possible geometrical isomers for the complex [Cr(en)2(CN)2]+.

Answer

3.4.7 https://chem.libretexts.org/@go/page/48468
Two geometrical isomers are possible: trans and cis.

Summary
The existence of coordination compounds with the same formula but different arrangements of the ligands was crucial in the
development of coordination chemistry. Two or more compounds with the same formula but different arrangements of the atoms
are called isomers. Many metal complexes form isomers, which are two or more compounds with the same formula but different
arrangements of atoms. Structural isomers differ in which atoms are bonded to one another, while geometrical isomers differ only
in the arrangement of ligands around the metal ion. Ligands adjacent to one another are cis, while ligands across from one another
are trans.
Table 3.4.1 emphasizes the key differences of the three classes of structural isomers discussed below. The highlighted ions are the
ions that switch or change somehow to make the type of structural isomer it is.
Table 3.4.1 : Overview of Structural Isomer Classes
Structural Isomer Examples

Ionization [CoBr(H2O)5]+Cl- and [CoCl(H2O)5]+Br-

Coordination [Zn(NH3)4]+[CuCl4]2- and [Cu(NH3)4]+[ZnCl4]2-

Linkage [Co(NO2)6]3- and [Co(ONO)6]3-

References
1. Petrucci, Harwood, Herring, and Madura. General Chemistry: Principles and Modern Applications: Ninth Edition. New Jersey:
Pearson, 2007.
2. Atkins, Peter, and Loretta Jones. Chemical Principles: The Quest for Insight: Fourth Edition. W. H. Freeman and Company,
2008. Print.

Contributors and Attributions


Prof. Robert J. Lancashire (The Department of Chemistry, University of the West Indies)
Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

3.4: Isomerism is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

3.4.8 https://chem.libretexts.org/@go/page/48468
3.5: Optical Isomerism in Metal Complexes
Optical isomers are stereoisomers that are related via non-superimposable mirror images of each other. They differ from geometric
stereoisomers in that they rotate the polarization of plane-polarized light. These isomers are referred to as enantiomers or chiral.

Introduction
Optical activity refers to whether or not a compound has optical isomers. A coordinate compound that is optically active has
optical isomers and a coordinate compound that is not optically active does not have optical isomers. As we will discuss later,
optical isomers have the unique property of rotating light. When light is shot through a polarimeter, optical isomers can rotate the
light so it comes out in with a different polarization on the other end. Armed with the knowledge of symmetry and mirror images,
optical isomers should not be very difficult. There are two ways optical isomers can be determined: using mirror images or using
planes of symmetry.
Optical isomers do not exhibit symmetry and do not have identical mirror images. Let's go through a quick review of symmetry and
mirror images. A mirror image of an object is that object flipped or the way the object would look in front of a mirror. For
example, the mirror image of your left hand would be your right hand. Symmetry on the other hand refers to when an object looks
exactly the same when sliced in a certain direction with a plane. For example imagine the shape of a square. No matter in what
direction it is sliced, the two resulting images will be the same.

What is a Polarimeter?
A polarimeter is a scientific instrument used to measure the angle of rotation caused by passing polarized light through an optically
active substance. Some chemical substances are optically active, and polarized (uni-directional) light will rotate either to the left
(counter-clockwise) or right (clockwise) when passed through these substances. The amount by which the light is rotated is known
as the angle of rotation. The polarimeter is made up of a polarizer (#3 on Figure 3.5.1) and an analyzer (#7 on Figure 3.5.1). The
polarizer allows only those light waves which move in a single plane. This causes the light to become plane polarized. When the
analyzer is also placed in a similar position it allows the light waves coming from the polarizer to pass through it. When it is rotated
through the right angle no waves can pass through the right angle and the field appears to be dark. If now a glass tube containing an
optically active solution is placed between the polarizer and analyzer the light now rotates through the plane of polarization through
a certain angle, the analyzer will have to be rotated in same angle.

Figure 3.5.1 : Schematic of a polarimeter showing the principles behind it's operation.The plane of polarization of plane
polarised light (4) rotates (6) as it passes through an optically active sample (5). This angle is determined with a rotatable polarizing
filter (7). (CC BY-SA 3.0; Kaidor via Wikipedia)
Chemists use polarimeters to investigate the influence of compounds (in the sample cell) on plane polarized light. Samples
composed only of achiral molecules (e.g. water or hexane), have no effect on the polarized light beam. However, if a single
enantiomer is examined (all sample molecules being right-handed, or all being left-handed), the plane of polarization is rotated in
either a clockwise (positive) or counter-clockwise (negative) direction, and the analyzer must be turned an appropriate matching
angle, α, if full light intensity is to reach the detector. In the above illustration, the sample has rotated the polarization plane
clockwise by +90º, and the analyzer has been turned this amount to permit maximum light transmission.

3.5.1 https://chem.libretexts.org/@go/page/81850
The observed rotations (α) of enantiomers are opposite in direction. One enantiomer will rotate polarized light in a clockwise
direction, termed dextrorotatory or (+), and its mirror-image partner in a counter-clockwise manner, termed levorotatory or (–).
The prefixes dextro and levo come from the Latin dexter, meaning right, and laevus, for left, and are abbreviated d and l
respectively. A 50:50 mixture of enantiomers has no observable optical activity. Such mixtures are called racemates or racemic
modifications.

The "Mirror Image Method" to Determine Optical Isomers


Optical isomers do not exhibit symmetry and do not have identical mirror images. Let's go through a quick review of symmetry and
mirror images. A mirror image of an object is that object flipped or the way the object would look in front of a mirror. For
example, the mirror image of your left hand would be your right hand. Symmetry on the other hand refers to when an object looks
exactly the same when sliced in a certain direction with a plane. For example imagine the shape of a square. No matter in what
direction it is sliced, the two resulting images will be the same.
The mirror image method uses a mirror image of the molecule to determine whether optical isomers exist or not. If the mirror
image can be rotated in such a way that it looks identical to the original molecule, then the molecule is said to be superimposable
and has no optical isomers. On the other hand, if the mirror image cannot be rotated in any way such that it looks identical to the
original molecule, then the molecule is said to be non-superimposable and the molecule has optical isomers. Once again, if the
mirror image is superimposable, then no optical isomers, but if the mirror image is non-superimposable, then optical isomers exist.
Non-superimposable means the structure cannot be rotated in a way that one can be put on top of another. This means that no
matter how the structure is rotated, it cannot be put on top of another with all points matching. An example of this is your hands.
Both left and right hands are identical, but they cannot be put on top of each other with all points matching.

Non-superimposable means the structure cannot be rotated in a way that one can be put
on top of another.

Example 3.5.1: The [OsCl 3


(CO) ]
3

Ion

Are there enantiomers for the f ac − [OsCl 3


(CO) ]
3

complex ion?
Solution
To use mirror image method, we draw the 3D structure for this geometric isomer (not the mer isomer) and then take a mirror
image like below.


The left image can be rotated by 180° through a vertical axis to superimpose on the orginal. So the mer-[OsCl 3
(CO) ]
3

complex ion is not optically active or we can call it achiral.

Exercise 3.5.1

Use the mirror image method to determine if there are enantiomers for the mer−[OsCl 3
(CO) ]
3

complex ion.

3.5.2 https://chem.libretexts.org/@go/page/81850
Example 3.5.2

Consider the octahedral complex with six monodentate ligands MA B C have both cis and trans isomers. However, the cis
2 2 2

isomer will also have two optical isomers (Figure 3.5.2). No matter how one rotates one of the mirror images, they are are not
superimposable.

Figure 3.5.2 : Two MA2


B C
2 2
enantiomers. (CC BY-SA-NC 4.0; ChemTube3D via Nick Greeves)

Exercise 3.5.1: MA 2
B C
2 2

Confirm to yourself that if two like ligands in MA B C


2 2 2
are trans to each other, then the complex will not have an optical
isomer.

Answer
If two like ligands are trans then there will be a mirror plane of symmetry normal to the line connecting the ligands that are
trans to each other. The presence of the mirror plane means this complex will not have an optical isomer.

The "Plane of Symmetry Method" to Determine Optical Isomers


The plane of symmetry method uses symmetry also to identify optical isomers. In this method, one tries to see if such a plane exists
which when cut through the coordinate compound produces two exact images. In other words, one looks for the existence of a
plane of symmetry within the coordinate compound. If a plane of symmetry exists, then no optical isomers exist for that complex
and it is called achiral.

Example 3.5.3: A Tetrahedral Complex

Consider the tetrahedral molecule, CHBrClF. Is this molecule optically active?

3.5.3 https://chem.libretexts.org/@go/page/81850
The CHBrClF molecule (note the color scheme: grey=carbon, white=hydrogen, green=chlorine, blue=fluorine, red=bromine)
Solution
Mirror-image method
First take the Mirror-image method. The mirror image of the molecule is:

Note that this mirror image is not superimposable. In other words, the mirror image above cannot be rotated in any such way
that it looks identical to the original molecule. Remember, if the mirror image is not superimposable, then optical isomers exist.
Thus we know that this molecule has optical isomers.
Plane of Symmetry Method
Let's try approaching this problem using the symmetry method. If we take the original molecule and draw an axis or plane of
symmetry down the middle, this is what we get:

Since the left side is not identical to the right, this molecule does not have a symmetrical center and thus can be called
chiral.Additionally, because it does not have a symmetrical center, we can conclude that this molecule has optical isomers. In
general, when dealing with a tetrahedral molecule that has 4 different ligands, optical isomers will exist most of the time.
No matter which method you use, the answer will end up being the same.

Optical isomers have no plane of symmetry. For tetrahedral complexes, this is fairly easy to recognize the possibility of this by
looking for a center atom with four different things attached to it. Unfortunately, this is not quite so easy with more complicated
geometries!

3.5.4 https://chem.libretexts.org/@go/page/81850
Example 3.5.4: An Octahedral Complex

Is the octahedral compound FeCl 3


F
3
optically active?

(note the color scheme: orange=iron, blue=fluorine, green=chlorine):


Solution
If we try to attempt this problem using the mirror image method, we notice that the mirror image is essentially identical to the
original molecule. In other words, the mirror image can be placed on top of the original molecule and is thus superimposable.
Since the mirror image is superimposable, this molecule does not have any optical isomers. Let's attempt this same problem
using the symmetry method. If we draw an axis or plane of symmetry, this is what we get:

Since the left side is identical to the right side, this molecule has a symmetrical center and is an achiral molecule. Thus, it has
no optical isomers.

Enantiomers in Octahedral Complexes with Bidentate Ligands


The examples you are most likely to need occur in octahedral complexes that contain bidentate ligands, e.g., complex ions like
2 + 3 −
[Ni (NH CH CH NH ) ]
2 2 2 2 3
or [Cr(C O ) ] . In bidentate complexes or chelating complexes, a ligand binds very tightly to
2 4 3

the metal because it holds onto the metal via more than one atom. Ethylenediamine is one example of a bidentate ligand.

Figure 3.5.3 : Ethylenediamine (sometimes abbreviated en) is an example of a bidentate ligand


As you may expect, bidentate ligands bind tightly to metals because they form two bonds with it, rather than just one. It is also
important to recognize that in an octahedral complex, the two donor atoms in most bidentate ligand bind cis to each other
and cannot reach all the way around the molecule to bind trans to each other (see below). The diagram below shows a simplified
view of one of the complex ions with three oxalato ligands (also a bidentate ligand), however, they all have the same shape - all
that differs is the nature of the "headphones".

3.5.5 https://chem.libretexts.org/@go/page/81850
A substance with no plane of symmetry is going to have optical isomers - one of which is the mirror image of the other. One of the
isomers will rotate the plane of polarization of plane polarized light clockwise; the other rotates it counter-clockwise. In this case,
the two isomers are:

Figure 3.5.4 : A left-handed propeller shape (left) is seen in one optical isomers of the M (ox) octahedral complex. The other
+
3

optical isomer is a right-handed propeller (right). These shapes are alternatively described as a left-handed screw and a right handed
screw. If you can picture turning the shape so that it screws into the page behind it, which direction would you turn the
screwdriver? If you would twist the screwdriver clockwise, then you have a right handed screw. If you would twist the screwdriver
counter-clockwise, then you have a left handed screw.

Most, but not all, Bidentate Ligands Bind to Metals in the cis Orientation

A wide variety of bidentate ligands bind to metals in the cis fashion, but few can bind in the trans geometry. These are called
trans-spanning ligands and required building a bridge between the bonding atoms (e.g. methylene brindges
−CH CH CH CH −).
2 2 2 2

Trans vs cis ligands in square planar complexes. (CC BY-SA 3.0; via Wikipedia)
We will not consider any trans-spanning ligands in this text.

A complex containing three bidentate ligands can take on the shape of a left-handed propeller or a right-handed propeller. You may
be able to see that there is no way of rotating the second isomer in space so that it looks exactly the same as the first one. As long as
you draw the isomers carefully, with the second one a true reflection of the first, the two structures will be different.
These kinds of complexes were historically important in demonstrating how small molecules and ions bound to metal cations. By
showing that some metal complexes were chiral and displayed optical activity, early 20th century workers such as Alfred Werner
were able to rule out some competing ideas about the structures of metal compounds. Today, we know that metal complexes play
important roles in enzymes in biology, and Werner's work on metal complexes laid the groundwork for how we think about these
complexes. In addition, stereochemistry in metal complexes became very important in the late 20th century, especially as
pharmaceutical companies looked for catalysts that could aid in the production of one enantiomer of a drug, and not the other, in
order to maximize pharmaceutical effectiveness and minimize side effects.

Exercise 3.5.1

Identify the relationships (types of isomers) between the following pairs of iron(III) oxalate ions.

3.5.6 https://chem.libretexts.org/@go/page/81850
3.5.7 https://chem.libretexts.org/@go/page/81850
Contributors and Attributions
Prof. Robert J. Lancashire (The Department of Chemistry, University of the West Indies)
Jim Clark (Chemguide.co.uk)
Chris P Schaller, Ph.D., (College of Saint Benedict / Saint John's University)

3.5: Optical Isomerism in Metal Complexes is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.

3.5.8 https://chem.libretexts.org/@go/page/81850
3.6: Stability Aspects of Complex-Ion Equilibria
Learning Objectives
To be introduced to complex ions, including ligands.

Previously, you learned that metal ions in aqueous solution are hydrated—that is, surrounded by a shell of usually four or six water
molecules. A hydrated ion is one kind of a complex ion (or, simply, complex), a species formed between a central metal ion and
one or more surrounding ligands, molecules or ions that contain at least one lone pair of electrons, such as the [Al(H2O)6]3+ ion.
A complex ion forms from a metal ion and a ligand because of a Lewis acid–base interaction. The positively charged metal ion acts
as a Lewis acid, and the ligand, with one or more lone pairs of electrons, acts as a Lewis base. Small, highly charged metal ions,
such as Cu2+ or Ru3+, have the greatest tendency to act as Lewis acids, and consequently, they have the greatest tendency to form
complex ions.

Figure 3.6.1 : The Formation of Complex Ions. An aqueous solution of CuSO4 consists of hydrated Cu2+ ions in the form of pale
blue [Cu(H2O)6]2+ (left). The addition of aqueous ammonia to the solution results in the formation of the intensely blue-violet
[Cu(NH3)4(H2O)2]2+ ions, usually written as [Cu(NH3)4]2+ ion (right) because ammonia, a stronger base than H2O, replaces water
molecules from the hydrated Cu2+ ion. For a more complete description, see https://www.youtube.com/watch?v=IQNcLH6OZK0.
As an example of the formation of complex ions, consider the addition of ammonia to an aqueous solution of the hydrated Cu2+ ion
{[Cu(H2O)6]2+}. Because it is a stronger base than H2O, ammonia replaces the water molecules in the hydrated ion to form the
[Cu(NH3)4(H2O)2]2+ ion. Formation of the [Cu(NH3)4(H2O)2]2+ complex is accompanied by a dramatic color change, as shown in
Figure 3.6.1. The solution changes from the light blue of [Cu(H2O)6]2+ to the blue-violet characteristic of the [Cu(NH3)4(H2O)2]2+
ion.

The Formation Constant


The replacement of water molecules from [Cu(H O) ] by ammonia occurs in sequential steps. Omitting the water molecules
2 6
2 +

bound to Cu for simplicity, we can write the equilibrium reactions as follows:


2 +

2 + 2 +
Cu (aq) + NH (aq) −
↽⇀
− [Cu(NH )] (aq) (step 1)
3 3

with
2 +
[Cu(NH )] (aq)
3
K1 =
2 +
[ Cu (aq)][ NH (aq)]
3

2 + 2 +
[Cu(NH )] (aq) + NH (aq) −
↽⇀
− [Cu(NH ) ] (aq) (step 2)
3 3 3 2

with
2 +
[Cu(NH ) ] (aq)
3 2
K2 =
2 +
[ [Cu(NH )] (aq)][ NH (aq)]
3 3

2 + 2 +
[Cu(NH ) ] (aq) + NH (aq) −
↽⇀
− [Cu(NH ) ] (aq) (step 3)
3 2 3 3 3

with
2 +
[Cu(NH ) ] (aq)
3 3
K3 =
2 +
[ [Cu(NH ) ] (aq)][ NH (aq)]
3 2 3

3.6.1 https://chem.libretexts.org/@go/page/48472
2 + 2 +
[Cu(NH ) ] (aq) + NH (aq) −
↽⇀
− [Cu(NH ) ] (aq) (step 4)
3 3 3 3 4

with
2 +
[Cu(NH ) ] (aq)
3 4
K4 =
2 +
[ [Cu(NH ) ] (aq)][ NH (aq)]
3 3 3

The sum of the stepwise reactions is the overall equation for the formation of the complex ion: The hydrated Cu2+ ion contains six
H2O ligands, but the complex ion that is produced contains only four N H ligands, not six. 3

2 + 2 +
Cu (aq) + 4 NH (aq) −
↽⇀
− [Cu(NH ) ] (aq) (3.6.1)
3 3 4

The equilibrium constant for the formation of the complex ion from the hydrated ion is called the formation constant (K ). The f

equilibrium constant expression for Kf has the same general form as any other equilibrium constant expression. In this case, the
expression is as follows:
2+
[Cu(NH3 )4 ] ]
Kf =
2+ 4
[C u ][NH3 ]

13
= 2.1 × 10

= K1 × K2 × K3 × K4

The formation constant (K ) has the same general form as any other equilibrium
f

constant expression.
Water, a pure liquid, does not appear explicitly in the equilibrium constant expression, and the hydrated Cu2+(aq) ion is represented
as Cu2+ for simplicity. As for any equilibrium, the larger the value of the equilibrium constant (in this case, Kf), the more stable the
product. With Kf = 2.1 × 1013, the [Cu(NH3)4(H2O)2]2+ complex ion is very stable. The formation constants for some common
complex ions are listed in Table 3.6.1.
Table 3.6.1 : Formation Constants for Selected Complex Ions in Aqueous Solution*
Complex Ion Equilibrium Equation Kf

[Ag(NH3)2]+ Ag+ + 2NH3 ⇌ [Ag(NH3)2]+ 1.1 × 107

Ammonia Complexes [Cu(NH3)4]2+ Cu2+ + 4NH3 ⇌ [Cu(NH3)4]2+ 2.1 × 1013

[Ni(NH3)6]2+ Ni2+ + 6NH3 ⇌ [Ni(NH3)6]2+ 5.5 × 108

[Ag(CN)2]− Ag+ + 2CN− ⇌ [Ag(CN)2]− 1.1 × 1018

Cyanide Complexes [Ni(CN)4]2− Ni2+ + 4CN− ⇌ [Ni(CN)4]2− 2.2 × 1031

[Fe(CN)6]3− Fe3+ + 6CN− ⇌ [Fe(CN)6]3− 1 × 1042

[Zn(OH)4]2− Zn2+ + 4OH− ⇌ [Zn(OH)4]2− 4.6 × 1017


Hydroxide Complexes
[Cr(OH)4]− Cr3+ + 4OH− ⇌ [Cr(OH)4]− 8.0 × 1029

[HgCl4]2− Hg2+ + 4Cl− ⇌ [HgCl4]2− 1.2 × 1015

Halide Complexes [CdI4]2− Cd2+ + 4I ⇌ [CdI4]2− 2.6 × 105

[AlF6]3− Al3+ + 6F− ⇌ [AlF6]3− 6.9 × 1019

[Ag(S2O3)2]3− Ag+ + 2S2O32− ⇌ [Ag(S2O3)2]3− 2.9 × 1013


Other Complexes
[Fe(C2O4)3]3− Fe3+ + 3C2O42− ⇌ [Fe(C2O4)3]3− 2.0 × 1020

*Reported values are overall formation constants. Source: Data from Lange’s Handbook of Chemistry, 15th ed.
(1999).

3.6.2 https://chem.libretexts.org/@go/page/48472
Example 3.6.1

If 12.5 g of Cu(NO 3
)
2
∙6H O
2
is added to 500 mL of 1.00 M aqueous ammonia, what is the equilibrium concentration of
(aq)?
2 +
Cu

Given: mass of Cu2+ salt and volume and concentration of ammonia solution
Asked for: equilibrium concentration of Cu2+(aq)
Strategy:
A. Calculate the initial concentration of Cu (aq) due to the addition of copper(II) nitrate hexahydrate. Use the stoichiometry
2 +

of the reaction shown in Equation 3.6.1 to construct a table showing the initial concentrations, the changes in
concentrations, and the final concentrations of all species in solution.
B. Substitute the final concentrations into the expression for the formation constant (K ) to calculate the equilibrium f

concentration of Cu (aq). 2 +

Solution
Adding an ionic compound that contains Cu (aq) to an aqueous ammonia solution will result in the formation of
2 +

[Cu(NH3)4]2+(aq), as shown in Equation 3.6.1 We assume that the volume change caused by adding solid copper(II) nitrate to
aqueous ammonia is negligible.
A The initial concentration of Cu 2 +
(aq) from the amount of added copper nitrate prior to any reaction is as follows:

1 mol 1 1000 mL
12.5 g Cu(NO3 )2 ⋅ 6 H2 O ( )( )( ) = 0.0846 M
295.65 g 500 mL 1 L

Because the stoichiometry of the reaction is four NH3 to one Cu2+, the amount of NH3 required to react completely with the
Cu2+ is 4(0.0846) = 0.338 M. The concentration of ammonia after complete reaction is 1.00 M − 0.338 M = 0.66 M. These
results are summarized in the first two lines of the following table. Because the equilibrium constant for the reaction is large
(2.1 × 1013), the equilibrium will lie far to the right. Thus we will assume that the formation of [Cu(NH3)4]2+ in the first step is
complete and allow some of it to dissociate into Cu2+ and NH3 until equilibrium has been reached. If we define x as the amount
of Cu2+ produced by the dissociation reaction, then the stoichiometry of the reaction tells us that the change in the
concentration of [Cu(NH3)4]2+ is −x, and the change in the concentration of ammonia is +4x, as indicated in the table. The final
concentrations of all species (in the bottom row of the table) are the sums of the concentrations after complete reaction and the
changes in concentrations.
2 + 2 +
Cu + 4 NH −
↽⇀
− [Cu(NH ) ]
3 3 4

ICE [Cu2+] [NH3] [[Cu(NH3)4]2+]

initial 0.0846 1.00 0

after complete reaction 0 0.66 0.0846

change +x +4x −x

final x 0.66 + 4x 0.0846 − x

B Substituting the final concentrations into the expression for the formation constant (K ) and assuming that f x ≪ 0.0846 ,
which allows us to remove x from the sum and difference,
2+
[[Cu(NH3 )4 ] ]
Kf =
2+ 4
[C u ][NH3 ]

0.0846 − x
=
x(0.66 + 4x)4

0.0846
13
≈ = 2.1 × 10
4
x(0.66)

−14
x = 2.1 × 10

3.6.3 https://chem.libretexts.org/@go/page/48472
The value of x indicates that our assumption was justified. The equilibrium concentration of Cu2+(aq) in a 1.00 M ammonia
solution is therefore 2.1 × 10−14 M.

Exercise 3.6.1

The ferrocyanide ion ([Fe(CN) ] ) is very stable, with a


6
4 −
Kf of 35
1 × 10 . Calculate the concentration of cyanide ion in
equilibrium with a 0.65 M solution of K [Fe(CN) ].
4 6

Answer
2 × 10−6 M

The Effect of the Formation of Complex Ions on Solubility


What happens to the solubility of a sparingly soluble salt if a ligand that forms a stable complex ion is added to the solution? One
such example occurs in conventional black-and-white photography. Recall that black-and-white photographic film contains light-
sensitive microcrystals of AgBr, or mixtures of AgBr and other silver halides. AgBr is a sparingly soluble salt, with a Ksp of 5.35 ×
10−13 at 25°C. When the shutter of the camera opens, the light from the object being photographed strikes some of the crystals on
the film and initiates a photochemical reaction that converts AgBr to black Ag metal. Well-formed, stable negative images appear
in tones of gray, corresponding to the number of grains of AgBr converted, with the areas exposed to the most light being darkest.
To fix the image and prevent more AgBr crystals from being converted to Ag metal during processing of the film, the unreacted
AgBr on the film is removed using a complexation reaction to dissolve the sparingly soluble salt.
The reaction for the dissolution of silver bromide is as follows:
+ −
AgBr(s) −
↽⇀
− Ag (aq) + Br (aq) (3.6.2)

with
−13
Ksp = 5.35 × 10  at 25°C (3.6.3)

+ − −7
The equilibrium lies far to the left, and the equilibrium concentrations of Ag and Br ions are very low (7.31 × 10 M). As a
result, removing unreacted AgBr from even a single roll of film using pure water would require tens of thousands of liters of water
and a great deal of time. Le Chatelier’s principle tells us, however, that we can drive the reaction to the right by removing one of
the products, which will cause more AgBr to dissolve. Bromide ion is difficult to remove chemically, but silver ion forms a variety
of stable two-coordinate complexes with neutral ligands, such as ammonia, or with anionic ligands, such as cyanide or thiosulfate
(S2O32−). In photographic processing, excess AgBr is dissolved using a concentrated solution of sodium thiosulfate.

The reaction of Ag+ with thiosulfate is as follows:


+ 2− 3−
Ag + 2 S2 O ⇌ [Ag(S2 O3 )2 ] (3.6.4)
(aq) 3(aq) (aq)

with
13
Kf = 2.9 × 10 (3.6.5)

The magnitude of the equilibrium constant indicates that almost all Ag+ ions in solution will be immediately complexed by
thiosulfate to form [Ag(S2O3)2]3−. We can see the effect of thiosulfate on the solubility of AgBr by writing the appropriate
reactions and adding them together:

3.6.4 https://chem.libretexts.org/@go/page/48472
+ − −13
AgBr(s) ⇌ Ag (aq) + Br (aq) Ksp = 5.35 × 10

+ 2− 3− 13
Ag (aq) + 2 S2 O (aq) ⇌ [Ag(S2 O3 )2 ] (aq) Kf = 2.9 × 10
3

2− 3− −
AgBr(s) + 2 S2 O (aq) ⇌ [Ag(S2 O3 )2 ] (aq) + Br (aq) K = Ksp Kf = 15
3

Comparing K with K shows that the formation of the complex ion increases the solubility of AgBr by approximately 3 × 1013.
sp

The dramatic increase in solubility combined with the low cost and the low toxicity explains why sodium thiosulfate is almost
universally used for developing black-and-white film. If desired, the silver can be recovered from the thiosulfate solution using any
of several methods and recycled.

If a complex ion has a large Kf, the formation of a complex ion can dramatically increase
the solubility of sparingly soluble salts.

Example 3.6.2: Common Ion Effect in Complexation

Due to the common ion effect, we might expect a salt such as AgCl to be much less soluble in a concentrated solution of KCl
than in water. Such an assumption would be incorrect, however, because it ignores the fact that silver ion tends to form a two-
coordinate complex with chloride ions (AgCl2−). Calculate the solubility of AgCl in each situation:
a. in pure water
b. in 1.0 M KCl solution, ignoring the formation of any complex ions
c. the same solution as in part (b) except taking the formation of complex ions into account, assuming that AgCl2− is the only
Ag+ complex that forms in significant concentrations
At 25°C, Ksp = 1.77 × 10−10 for AgCl and Kf = 1.1 × 105 for AgCl2−.
Given: Ksp of AgCl, Kf of AgCl2−, and KCl concentration
Asked for: solubility of AgCl in water and in KCl solution with and without the formation of complex ions
Strategy:
A. Write the solubility product expression for AgCl and calculate the concentration of Ag+ and Cl− in water.
B. Calculate the concentration of Ag+ in the KCl solution.
C. Write balanced chemical equations for the dissolution of AgCl and for the formation of the AgCl2− complex. Add the two
equations and calculate the equilibrium constant for the overall equilibrium.
D. Write the equilibrium constant expression for the overall reaction. Solve for the concentration of the complex ion.
Solution
A. If we let x equal the solubility of AgCl, then at equilibrium
+ −
[ Ag ]=[ Cl ]=x M.

Substituting this value into the solubility product expression,


+ −
Ksp = [ Ag ][ Cl ]

= (x)(x)

2
=x

−10
= 1.77 × 10

−5
x = 1.33 × 10

Thus the solubility of AgCl in pure water at 25°C is 1.33 × 10−5 M.


B. If x equals the solubility of AgCl in the KCl solution, then at equilibrium [Ag+] = x M and [Cl−] = (1.0 + x) M. Substituting
these values into the solubility product expression and assuming that x ≤ 1.0:

3.6.5 https://chem.libretexts.org/@go/page/48472
+ −
Ksp = [ Ag ][ Cl ]

= (x)(1.0 + x)

≈ x(1.0)

−10
= 1.77 × 10 =x

If the common ion effect were the only important factor, we would predict that AgCl is approximately five orders of
magnitude less soluble in a 1.0 M KCl solution than in water.
C. To account for the effects of the formation of complex ions, we must first write the equilibrium equations for both the
dissolution and the formation of complex ions. Adding the equations corresponding to Ksp and Kf gives us an equation that
describes the dissolution of AgCl in a KCl solution. The equilibrium constant for the reaction is therefore the product of Ksp
and Kf:
+ − −10
AgCl(s) ⇌ Ag (aq) + C l (aq) Ksp = 1.77 × 10

+ − − 5
Ag aq) + 2C l ⇌ [AgC l2 ] Kf = 1.1 × 10

− − −5
AgCl(s) + C l ⇌ [AgC l2 ] K = Ksp Kf = 1.9 × 10

D. If we let x equal the solubility of AgCl in the KCl solution, then at equilibrium [AgCl2−] = x and [Cl−] = 1.0 − x.
Substituting these quantities into the equilibrium constant expression for the net reaction and assuming that x ≤ 1.0,

[AgC l ]
2
K =

[C l ]

x
−5
= ≈ 1.9 × 10
1.0 − x

=x

That is, AgCl dissolves in 1.0 M KCl to produce a 1.9 × 10−5 M solution of the AgCl2− complex ion. Thus we predict that
AgCl has approximately the same solubility in a 1.0 M KCl solution as it does in pure water, which is 105 times greater than
that predicted based on the common ion effect. (In fact, the measured solubility of AgCl in 1.0 M KCl is almost a factor of 10
greater than that in pure water, largely due to the formation of other chloride-containing complexes.)

Exercise 3.6.2

Calculate the solubility of mercury(II) iodide (HgI ) in each situation:


2

a. pure water
b. a 3.0 M solution of NaI, assuming [HgI4]2− is the only Hg-containing species present in significant amounts
Ksp = 2.9 × 10−29 for HgI2 and Kf = 6.8 × 1029 for [HgI4]2−.

Answer a
1.9 × 10−10 M
Answer b
1.4 M

Complexing agents, molecules or ions that increase the solubility of metal salts by forming soluble metal complexes, are common
components of laundry detergents. Long-chain carboxylic acids, the major components of soaps, form insoluble salts with Ca2+ and
Mg2+, which are present in high concentrations in “hard” water. The precipitation of these salts produces a bathtub ring and gives a
gray tinge to clothing.

3.6.6 https://chem.libretexts.org/@go/page/48472
Adding a complexing agent such as pyrophosphate (O3POPO34−, or P2O74−) or triphosphate (P3O105−) to detergents prevents the
magnesium and calcium salts from precipitating because the equilibrium constant for complex-ion formation is large:

2 + 4 − 2 −
Ca (aq) + O POPO (aq) −
↽⇀
− [Ca(O POPO )] (aq)
3 4 3 3

with
4
Kf = 4 × 10

However, phosphates can cause environmental damage by promoting eutrophication, the growth of excessive amounts of algae in a
body of water, which can eventually lead to large decreases in levels of dissolved oxygen that kill fish and other aquatic organisms.
Consequently, many states in the United States have banned the use of phosphate-containing detergents, and France has banned
their use beginning in 2007. “Phosphate-free” detergents contain different kinds of complexing agents, such as derivatives of acetic
acid or other carboxylic acids. The development of phosphate substitutes is an area of intense research.

Commercial water softeners also use a complexing agent to treat hard water by passing the water over ion-exchange resins, which
are complex sodium salts. When water flows over the resin, sodium ion is dissolved, and insoluble salts precipitate onto the resin
surface. Water treated in this way has a saltier taste due to the presence of Na+, but it contains fewer dissolved minerals.

Complexing Agents in MRIs


Another application of complexing agents is found in medicine. Unlike x-rays, magnetic resonance imaging (MRI) can give
relatively good images of soft tissues such as internal organs. MRI is based on the magnetic properties of the 1H nucleus of
hydrogen atoms in water, which is a major component of soft tissues. Because the properties of water do not depend very much
on whether it is inside a cell or in the blood, it is hard to get detailed images of these tissues that have good contrast. To solve
this problem, scientists have developed a class of metal complexes known as “MRI contrast agents.” Injecting an MRI contrast
agent into a patient selectively affects the magnetic properties of water in cells of normal tissues, in tumors, or in blood vessels
and allows doctors to “see” each of these separately (Figure 3.6.2). One of the most important metal ions for this application is
Gd3+, which with seven unpaired electrons is highly paramagnetic. Because Gd3+(aq) is quite toxic, it must be administered as
a very stable complex that does not dissociate in the body and can be excreted intact by the kidneys. The complexing agents
used for gadolinium are ligands such as DTPA5− (diethylene triamine pentaacetic acid), whose fully protonated form is shown
in Figure 3.6.2.

3.6.7 https://chem.libretexts.org/@go/page/48472
Figure 3.6.2 : An MRI Image of the Heart, Arteries, and Veins. When a patient is injected with a paramagnetic metal cation in
the form of a stable complex known as an MRI contrast agent, the magnetic properties of water in cells are altered. Because the
different environments in different types of cells respond differently, a physician can obtain detailed images of soft tissues.

Summary
The formation of complex ions can substantially increase the solubility of sparingly soluble salts if the complex ion has a large Kf.
A complex ion is a species formed between a central metal ion and one or more surrounding ligands, molecules or ions that contain
at least one lone pair of electrons. Small, highly charged metal ions have the greatest tendency to act as Lewis acids and form
complex ions. The equilibrium constant for the formation of the complex ion is the formation constant (Kf). The formation of a
complex ion by adding a complexing agent increases the solubility of a compound.

3.6: Stability Aspects of Complex-Ion Equilibria is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.
24.8: Aspects of Complex-Ion Equilibria is licensed CC BY-NC-SA 4.0.

3.6.8 https://chem.libretexts.org/@go/page/48472
3.7: Chelation
Ligands like chloride, water, and ammonia are said to be monodentate (one-toothed, from the Greek mono, meaning “one,” and the
Latin dent-, meaning “tooth”): they are attached to the metal via only a single atom. Ligands can, however, be bidentate (two-
toothed, from the Greek di, meaning “two”), tridentate (three-toothed, from the Greek tri, meaning “three”), or, in general,
polydentate (many-toothed, from the Greek poly, meaning “many”), indicating that they are attached to the metal at two, three, or
several sites, respectively. Ethylenediamine (H2NCH2CH2NH2, often abbreviated as en) and diethylenetriamine
(H2NCH2CH2NHCH2CH2NH2, often abbreviated as dien) are examples of a bidentate and a tridentate ligand, respectively, because
each nitrogen atom has a lone pair that can be shared with a metal ion. When a bidentate ligand such as ethylenediamine binds to a
metal such as Ni2+, a five-membered ring is formed. A metal-containing ring like that shown is called a chelate ring (from the
Greek chele, meaning “claw”). Correspondingly, a polydentate ligand is a chelating agent, and complexes that contain polydentate
ligands are called chelate complexes.

Experimentally, it is observed that metal complexes of polydentate ligands are significantly more stable than the corresponding
complexes of chemically similar monodentate ligands; this increase in stability is called the chelate effect. For example, the
complex of Ni2+ with three ethylenediamine ligands, [Ni(en)3]2+, should be chemically similar to the Ni2+ complex with six
ammonia ligands, [Ni(NH3)6]2+. In fact, the equilibrium constant for the formation of [Ni(en)3]2+ is almost 10 orders of magnitude
larger than the equilibrium constant for the formation of [Ni(NH3)6]2+ (Table E4):
2+ 2+ 8
[Ni(H2 O)6 ] + 6NH3 ⇌ [Ni(NH3 )6 ] + 6 H2 O(l) Kf = 4 × 10 (3.7.1)

2+ 2+ 18
[Ni(H2 O)6 ] + 3en ⇌ [Ni(en)3 ] + 6 H2 O(l) Kf = 2 × 10 (3.7.2)

The formation constants are formulated as ligand exchange reactions with aqua ligands being displaced by new ligands (N H or 3

en ) in the examples above.

Chelate complexes are more stable than the analogous complexes with monodentate
ligands.
The stability of a chelate complex depends on the size of the chelate rings. For ligands with a flexible organic backbone like
ethylenediamine, complexes that contain five-membered chelate rings, which have almost no strain, are significantly more stable
than complexes with six-membered chelate rings, which are in turn much more stable than complexes with four- or seven-
membered rings. For example, the complex of nickel (II) with three ethylenediamine ligands is about 363,000 times more stable
than the corresponding nickel (II) complex with trimethylenediamine (H2NCH2CH2CH2NH2, abbreviated as tn):
2+ 2+ 17
[Ni(H2 O)6 ] + 3en ⇌ [Ni(en)3 ] + 6 H2 O(l) Kf = 6.76 × 10 (3.7.3)

2+ 2+ 12
[Ni(H2 O)6 ] + 3tn ⇌ [Ni(tn)3 ] + 6 H2 O(l) Kf = 1.86 × 10 (3.7.4)

*The above measurements were done in a solution of ionic strength 0.15 at 25º C.

Example 3.7.1

Arrange [Cr(en)3]3+, [CrCl6]3−, [CrF6]3−, and [Cr(NH3)6]3+ in order of increasing stability.


Given: four Cr(III) complexes
Asked for: relative stabilities
Strategy:
A Determine the relative basicity of the ligands to identify the most stable complexes.
B Decide whether any complexes are further stabilized by a chelate effect and arrange the complexes in order of increasing
stability.
Solution

3.7.1 https://chem.libretexts.org/@go/page/81584
A The metal ion is the same in each case: Cr3+. Consequently, we must focus on the properties of the ligands to determine the
stabilities of the complexes. Because the stability of a metal complex increases as the basicity of the ligands increases, we need
to determine the relative basicity of the four ligands. Our earlier discussion of acid–base properties suggests that ammonia and
ethylenediamine, with nitrogen donor atoms, are the most basic ligands. The fluoride ion is a stronger base (it has a higher
charge-to-radius ratio) than chloride, so the order of stability expected due to ligand basicity is
[CrCl6]3− < [CrF6]3− < [Cr(NH3)6]3+ ≈ [Cr(en)3]3+.
B Because of the chelate effect, we expect ethylenediamine to form a stronger complex with Cr3+ than ammonia.
Consequently, the likely order of increasing stability is
[CrCl6]3− < [CrF6]3− < [Cr(NH3)6]3+ < [Cr(en)3]3+.

Exercise 3.7.1

Arrange [Co(NH3)6]3+, [CoF6]3−, and [Co(en)3]3+ in order of decreasing stability.


Answer: [Co(en)3]3+ > [Co(NH3)6]3+ > [CoF6]3−

The Chelate Effect


The chelate effect can be seen by comparing the reaction of a chelating ligand and a metal ion with the corresponding reaction
involving comparable monodentate ligands. For example, comparison of the binding of 2,2'-bipyridine with pyridine or 1,2-
diaminoethane (ethylenediamine=en) with ammonia. It has been known for many years that a comparison of this type always
shows that the complex resulting from coordination with the chelating ligand is much more thermodynamically stable. This can
be seen by looking at the values for adding two monodentates compared with adding one bidentate, or adding four monodentates
compared to two bidentates, or adding six monodentates compared to three bidentates.

The Chelate Effect is that complexes resulting from coordination with the chelating ligand
is much more thermodynamically stable than complexes with non-chelating ligands.
A number of points should be highlighted from the formation constants in Table E4. In Table 3.7.1, it can be seen that the ΔH°
values for the formation steps are almost identical, that is, heat is evolved to about the same extent whether forming a complex
involving monodentate ligands or bidentate ligands. What is seen to vary significantly is the ΔS° term which changes from negative
(unfavorable) to positive (favorable). Note as well that there is a dramatic increase in the size of the ΔS° term for adding two
compared to adding four monodentate ligands. (-5 to -35 JK-1mol-1). What does this imply, if we consider ΔS° to give a measure of
disorder?
In the case of complex formation of Ni2+ with ammonia or 1,2-diaminoethane, by rewriting the equilibria, the following equations
are produced.

Using the equilibrium constant for the reaction (3 above) where the three bidentate ligands replace the six monodentateligands, we
find that at a temperature of 25° C:

ΔG = −2.303 RT log K (3.7.5)
10

= −2.303 RT (18.28 − 8.61) (3.7.6)

−1
= −54 kJ mol (3.7.7)

Based on measurements made over a range of temperatures, it is possible to break down the ΔG

term into the enthalpy and
entropy components.
∘ ∘ ∘
ΔG = ΔH − T ΔS (3.7.8)

3.7.2 https://chem.libretexts.org/@go/page/81584
The result is that: ΔH ∘
= −29kJmol
−1

- TΔS° = -25 kJ mol-1


and at 25C (298K)
ΔS° = +88 J K-1 mol-1

For many years, these numbers have been incorrectly recorded in textbooks. For example, the third edition of "Basic
Inorganic Chemistry" by F.A. Cotton, G. Wilkinson and P.L. Gaus, John Wiley & Sons, Inc, 1995, on page 186 gives the values
as:
ΔG° = -67 kJ mol-1
ΔH° = -12 kJ mol-1
-TΔS° = -55 kJ mol-1
The conclusion they drew from these incorrect numbers was that the chelate effect was essentially an entropy effect, since the
TΔS° contribution was nearly 5 times bigger than ΔH°.
In fact, the breakdown of the ΔG° into ΔH° and TΔS° shows that the two terms are nearly equal (-29 and -25 kJ mol-1,
respectively) with the ΔH° term a little bigger! The entropy term found is still much larger than for reactions involving a non-
chelating ligand substitution at a metal ion. How can we explain this enhanced contribution from entropy? One explanation is
to count the number of species on the left and right hand side of the equation above.

It will be seen that on the left-hand-side there are 4 species, whereas on the right-hand-side there are 7 species, that is a net gain of
3 species occurs as the reaction proceeds. This can account for the increase in entropy since it represents an increase in the disorder
of the system. An alternative view comes from trying to understand how the reactions might proceed. To form a complex with 6
monodentates requires 6 separate favorable collisions between the metal ion and the ligand molecules. To form the tris-bidentate
metal complex requires an initial collision for the first ligand to attach by one arm but remember that the other arm is always going
to be nearby and only requires a rotation of the other end to enable the ligand to form the chelate ring.
If you consider dissociation steps, then when a monodentate group is displaced, it is lost into the bulk of the solution. On the other
hand, if one end of a bidentate group is displaced the other arm is still attached and it is only a matter of the arm rotating around
and it can be reattached again. Both conditions favor the formation of the complex with bidentate groups rather than monodentate
groups.

3.7: Chelation is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

3.7.3 https://chem.libretexts.org/@go/page/81584
3.8: Bonding in Octahedral Complex Ions- Crystal Field Theory
Learning Objectives
To understand how crystal field theory explains the electronic structures and colors of metal complexes.

One of the most striking characteristics of transition-metal complexes is the wide range of colors they exhibit. In this section, we
describe crystal field theory (CFT), a bonding model that explains many important properties of transition-metal complexes,
including their colors, magnetism, structures, stability, and reactivity. The central assumption of CFT is that metal–ligand
interactions are purely electrostatic in nature. Even though this assumption is clearly not valid for many complexes, such as those
that contain neutral ligands like CO, CFT enables chemists to explain many of the properties of transition-metal complexes with a
reasonable degree of accuracy. The Learning Objective of this Module is to understand how crystal field theory explains the
electronic structures and colors of metal complexes.

d-Orbital Splittings
CFT focuses on the interaction of the five (n − 1)d orbitals with ligands arranged in a regular array around a transition-metal ion.
We will focus on the application of CFT to octahedral complexes, which are by far the most common and the easiest to visualize.
Other common structures, such as square planar complexes, can be treated as a distortion of the octahedral model. According to
CFT, an octahedral metal complex forms because of the electrostatic interaction of a positively charged metal ion with six
negatively charged ligands or with the negative ends of dipoles associated with the six ligands. In addition, the ligands interact with
one other electrostatically. As you learned in our discussion of the valence-shell electron-pair repulsion (VSEPR) model, the
lowest-energy arrangement of six identical negative charges is an octahedron, which minimizes repulsive interactions between the
ligands.
We begin by considering how the energies of the d orbitals of a transition-metal ion are affected by an octahedral arrangement of
six negative charges. Recall that the five d orbitals are initially degenerate (have the same energy). If we distribute six negative
charges uniformly over the surface of a sphere, the d orbitals remain degenerate, but their energy will be higher due to repulsive
electrostatic interactions between the spherical shell of negative charge and electrons in the d orbitals (Figure 3.8.1a). Placing the
six negative charges at the vertices of an octahedron does not change the average energy of the d orbitals, but it does remove their
degeneracy: the five d orbitals split into two groups whose energies depend on their orientations. As shown in Figure 3.8.1b, the dz2
and dx2−y2 orbitals point directly at the six negative charges located on the x, y, and z axes. Consequently, the energy of an electron
in these two orbitals (collectively labeled the eg orbitals) will be greater than it will be for a spherical distribution of negative
charge because of increased electrostatic repulsions. In contrast, the other three d orbitals (dxy, dxz, and dyz, collectively called the
t2g orbitals) are all oriented at a 45° angle to the coordinate axes, so they point between the six negative charges. The energy of an
electron in any of these three orbitals is lower than the energy for a spherical distribution of negative charge.

3.8.1 https://chem.libretexts.org/@go/page/48469
Figure 3.8.1 : An Octahedral Arrangement of Six Negative Charges around a Metal Ion Causes the Five d Orbitals to Split into Two
Sets with Different Energies. (a) Distributing a charge of −6 uniformly over a spherical surface surrounding a metal ion causes the
energy of all five d orbitals to increase due to electrostatic repulsions, but the five d orbitals remain degenerate. Placing a charge of
−1 at each vertex of an octahedron causes the d orbitals to split into two groups with different energies: the dx2−y2 and dz2 orbitals
increase in energy, while the dxy, dxz, and dyz orbitals decrease in energy. The average energy of the five d orbitals is the same as for
a spherical distribution of a −6 charge, however. Attractive electrostatic interactions between the negatively charged ligands and the
positively charged metal ion (far right) cause all five d orbitals to decrease in energy but does not affect the splittings of the
orbitals. (b) The two eg orbitals (left) point directly at the six negatively charged ligands, which increases their energy compared
with a spherical distribution of negative charge. In contrast, the three t2g orbitals (right) point between the negatively charged
ligands, which decreases their energy compared with a spherical distribution of charge. (CC BY-SA-NC; anonymous by request)
The difference in energy between the two sets of d orbitals is called the crystal field splitting energy (Δo), where the subscript o
stands for octahedral. As we shall see, the magnitude of the splitting depends on the charge on the metal ion, the position of the
metal in the periodic table, and the nature of the ligands. (Crystal field splitting energy also applies to tetrahedral complexes: Δt.) It
is important to note that the splitting of the d orbitals in a crystal field does not change the total energy of the five d orbitals: the
two eg orbitals increase in energy by 0.6Δo, whereas the three t2g orbitals decrease in energy by 0.4Δo. Thus the total change in
energy is

2(0.6 Δo ) + 3(−0.4 Δo ) = 0. (3.8.1)

Crystal field splitting does not change the total energy of the d orbitals.
Thus far, we have considered only the effect of repulsive electrostatic interactions between electrons in the d orbitals and the six
negatively charged ligands, which increases the total energy of the system and splits the d orbitals. Interactions between the
positively charged metal ion and the ligands results in a net stabilization of the system, which decreases the energy of all five d
orbitals without affecting their splitting (as shown at the far right in Figure 3.8.1a).

3.8.2 https://chem.libretexts.org/@go/page/48469
Electronic Structures of Metal Complexes
We can use the d-orbital energy-level diagram in Figure 3.8.1 to predict electronic structures and some of the properties of
transition-metal complexes. We start with the Ti3+ ion, which contains a single d electron, and proceed across the first row of the
transition metals by adding a single electron at a time. We place additional electrons in the lowest-energy orbital available, while
keeping their spins parallel as required by Hund’s rule. As shown in Figure 3.8.2, for d1–d3 systems—such as [Ti(H2O)6]3+,
[V(H2O)6]3+, and [Cr(H2O)6]3+, respectively—the electrons successively occupy the three degenerate t2g orbitals with their spins
parallel, giving one, two, and three unpaired electrons, respectively. We can summarize this for the complex [Cr(H2O)6]3+, for
example, by saying that the chromium ion has a d3 electron configuration or, more succinctly, Cr3+ is a d3 ion.

Figure 3.8.2 : The Possible Electron Configurations for Octahedral dn Transition-Metal Complexes (n = 1–10). Two different
configurations are possible for octahedral complexes of metals with d4, d5, d6, and d7 configurations; the magnitude of Δo
determines which configuration is observed. (CC BY-SA-NC; anonymous by request)
When we reach the d4 configuration, there are two possible choices for the fourth electron: it can occupy either one of the empty eg
orbitals or one of the singly occupied t2g orbitals. Recall that placing an electron in an already occupied orbital results in
electrostatic repulsions that increase the energy of the system; this increase in energy is called the spin-pairing energy (P). If Δo is
less than P, then the lowest-energy arrangement has the fourth electron in one of the empty eg orbitals. Because this arrangement
results in four unpaired electrons, it is called a high-spin configuration, and a complex with this electron configuration, such as the
[Cr(H2O)6]2+ ion, is called a high-spin complex. Conversely, if Δo is greater than P, then the lowest-energy arrangement has the
fourth electron in one of the occupied t2g orbitals. Because this arrangement results in only two unpaired electrons, it is called a
low-spin configuration, and a complex with this electron configuration, such as the [Mn(CN)6]3− ion, is called a low-spin complex.
Similarly, metal ions with the d5, d6, or d7 electron configurations can be either high spin or low spin, depending on the magnitude
of Δo.
In contrast, only one arrangement of d electrons is possible for metal ions with d8–d10 electron configurations. For example, the
[Ni(H2O)6]2+ ion is d8 with two unpaired electrons, the [Cu(H2O)6]2+ ion is d9 with one unpaired electron, and the [Zn(H2O)6]2+
ion is d10 with no unpaired electrons.

If Δo is less than the spin-pairing energy, a high-spin configuration results. Conversely, if


Δo is greater, a low-spin configuration forms.

Factors That Affect the Magnitude of Δo


The magnitude of Δo dictates whether a complex with four, five, six, or seven d electrons is high spin or low spin, which affects its
magnetic properties, structure, and reactivity. Large values of Δo (i.e., Δo > P) yield a low-spin complex, whereas small values of
Δo (i.e., Δo < P) produce a high-spin complex. As we noted, the magnitude of Δo depends on three factors: the charge on the metal
ion, the principal quantum number of the metal (and thus its location in the periodic table), and the nature of the ligand. Values of
Δo for some representative transition-metal complexes are given in Table 3.8.1.
Table 3.8.1 : Crystal Field Splitting Energies for Some Octahedral (Δo)* and Tetrahedral (Δt) Transition-Metal Complexes
Octahedral Tetrahedral
Δo (cm−1) Octahedral Complexes Δo (cm−1) Δt (cm−1)
Complexes Complexes

*Energies obtained by spectroscopic measurements are often given in units of wave numbers (cm−1); the wave number is the reciprocal of the
wavelength of the corresponding electromagnetic radiation expressed in centimeters: 1 cm−1 = 11.96 J/mol.

3.8.3 https://chem.libretexts.org/@go/page/48469
Octahedral Tetrahedral
Δo (cm−1) Octahedral Complexes Δo (cm−1) Δt (cm−1)
Complexes Complexes

[Ti(H2O)6]3+ 20,300 [Fe(CN)6]4− 32,800 VCl4 9010

[V(H2O)6]2+ 12,600 [Fe(CN)6]3− 35,000 [CoCl4]2− 3300

[V(H2O)6]3+ 18,900 [CoF6]3− 13,000 [CoBr4]2− 2900

[CrCl6]3− 13,000 [Co(H2O)6]2+ 9300 [CoI4]2− 2700

[Cr(H2O)6]2+ 13,900 [Co(H2O)6]3+ 27,000

[Cr(H2O)6]3+ 17,400 [Co(NH3)6]3+ 22,900

[Cr(NH3)6]3+ 21,500 [Co(CN)6]3− 34,800

[Cr(CN)6]3− 26,600 [Ni(H2O)6]2+ 8500

Cr(CO)6 34,150 [Ni(NH3)6]2+ 10,800

[MnCl6]4− 7500 [RhCl6]3− 20,400

[Mn(H2O)6]2+ 8500 [Rh(H2O)6]3+ 27,000

[MnCl6]3− 20,000 [Rh(NH3)6]3+ 34,000

[Mn(H2O)6]3+ 21,000 [Rh(CN)6]3− 45,500

[Fe(H2O)6]2+ 10,400 [IrCl6]3− 25,000

[Fe(H2O)6]3+ 14,300 [Ir(NH3)6]3+ 41,000

*Energies obtained by spectroscopic measurements are often given in units of wave numbers (cm−1); the wave number is the reciprocal of the
wavelength of the corresponding electromagnetic radiation expressed in centimeters: 1 cm−1 = 11.96 J/mol.

Source of data: Duward F. Shriver, Peter W. Atkins, and Cooper H. Langford, Inorganic Chemistry, 2nd ed. (New York: W. H.
Freeman and Company, 1994).

Factor 1: Charge on the Metal Ion


Increasing the charge on a metal ion has two effects: the radius of the metal ion decreases, and negatively charged ligands are more
strongly attracted to it. Both factors decrease the metal–ligand distance, which in turn causes the negatively charged ligands to
interact more strongly with the d orbitals. Consequently, the magnitude of Δo increases as the charge on the metal ion increases.
Typically, Δo for a tripositive ion is about 50% greater than for the dipositive ion of the same metal; for example, for [V(H2O)6]2+,
Δo = 11,800 cm−1; for [V(H2O)6]3+, Δo = 17,850 cm−1.

Factor 2: Principal Quantum Number of the Metal


For a series of complexes of metals from the same group in the periodic table with the same charge and the same ligands, the
magnitude of Δo increases with increasing principal quantum number: Δo (3d) < Δo (4d) < Δo (5d). The data for hexaammine
complexes of the trivalent Group 9 metals illustrate this point:
[Co(NH3)6]3+: Δo = 22,900 cm−1
[Rh(NH3)6]3+: Δo = 34,100 cm−1
[Ir(NH3)6]3+: Δo = 40,000 cm−1
The increase in Δo with increasing principal quantum number is due to the larger radius of valence orbitals down a column. In
addition, repulsive ligand–ligand interactions are most important for smaller metal ions. Relatively speaking, this results in shorter
M–L distances and stronger d orbital–ligand interactions.

Factor 3: The Nature of the Ligands


Experimentally, it is found that the Δo observed for a series of complexes of the same metal ion depends strongly on the nature of
the ligands. For a series of chemically similar ligands, the magnitude of Δo decreases as the size of the donor atom increases. For
example, Δo values for halide complexes generally decrease in the order F− > Cl− > Br− > I− because smaller, more localized
charges, such as we see for F−, interact more strongly with the d orbitals of the metal ion. In addition, a small neutral ligand with a

3.8.4 https://chem.libretexts.org/@go/page/48469
highly localized lone pair, such as NH3, results in significantly larger Δo values than might be expected. Because the lone pair
points directly at the metal ion, the electron density along the M–L axis is greater than for a spherical anion such as F−. The
experimentally observed order of the crystal field splitting energies produced by different ligands is called the spectrochemical
series, shown here in order of decreasing Δo:
− − − 2− − − − − −
CO ≈ C N > NO > en > NH3 > SC N > H2 O > oxalate > OH > F > acetate > Cl > Br >I
2
strong-field ligands intermediate-field ligands weak-field ligands

The values of Δo listed in Table 3.8.1 illustrate the effects of the charge on the metal ion, the principal quantum number of the
metal, and the nature of the ligand.

The largest Δo splittings are found in complexes of metal ions from the third row of the
transition metals with charges of at least +3 and ligands with localized lone pairs of
electrons.

Example 3.8.1

For each complex, predict its structure, whether it is high spin or low spin, and the number of unpaired electrons present.
a. [CoF6]3−
b. [Rh(CO)2Cl2]−
Given: complexes
Asked for: structure, high spin versus low spin, and the number of unpaired electrons
Strategy:
a. From the number of ligands, determine the coordination number of the compound.
b. Classify the ligands as either strong field or weak field and determine the electron configuration of the metal ion.
c. Predict the relative magnitude of Δo and decide whether the compound is high spin or low spin.
d. Place the appropriate number of electrons in the d orbitals and determine the number of unpaired electrons.
Solution
a. A With six ligands, we expect this complex to be octahedral.
B The fluoride ion is a small anion with a concentrated negative charge, but compared with ligands with localized lone
pairs of electrons, it is weak field. The charge on the metal ion is +3, giving a d6 electron configuration.
C Because of the weak-field ligands, we expect a relatively small Δo, making the compound high spin.
D In a high-spin octahedral d6 complex, the first five electrons are placed individually in each of the d orbitals with their
spins parallel, and the sixth electron is paired in one of the t2g orbitals, giving four unpaired electrons.
b. A This complex has four ligands, so it is either square planar or tetrahedral.
B C Because rhodium is a second-row transition metal ion with a d8 electron configuration and CO is a strong-field
ligand, the complex is likely to be square planar with a large Δo, making it low spin. Because the strongest d-orbital
interactions are along the x and y axes, the orbital energies increase in the order dz2dyz, and dxz (these are degenerate); dxy;
and dx2−y2.
D The eight electrons occupy the first four of these orbitals, leaving the dx2−y2. orbital empty. Thus there are no unpaired
electrons.

Exercise 3.8.1

For each complex, predict its structure, whether it is high spin or low spin, and the number of unpaired electrons present.
a. [Mn(H2O)6]2+
b. [PtCl4]2−

Answer a

3.8.5 https://chem.libretexts.org/@go/page/48469
octahedral; high spin; five
Answer b
square planar; low spin; no unpaired electrons

Crystal Field Stabilization Energies


Recall that stable molecules contain more electrons in the lower-energy (bonding) molecular orbitals in a molecular orbital diagram
than in the higher-energy (antibonding) molecular orbitals. If the lower-energy set of d orbitals (the t2g orbitals) is selectively
populated by electrons, then the stability of the complex increases. For example, the single d electron in a d1 complex such as
[Ti(H2O)6]3+ is located in one of the t2g orbitals. Consequently, this complex will be more stable than expected on purely
electrostatic grounds by 0.4Δo. The additional stabilization of a metal complex by selective population of the lower-energy d
orbitals is called its crystal field stabilization energy (CFSE). The CFSE of a complex can be calculated by multiplying the number
of electrons in t2g orbitals by the energy of those orbitals (−0.4Δo), multiplying the number of electrons in eg orbitals by the energy
of those orbitals (+0.6Δo), and summing the two. Table 3.8.2 gives CFSE values for octahedral complexes with different d electron
configurations. The CFSE is highest for low-spin d6 complexes, which accounts in part for the extraordinarily large number of
Co(III) complexes known. The other low-spin configurations also have high CFSEs, as does the d3 configuration.
Table 3.8.2 : CFSEs for Octahedral Complexes with Different Electron Configurations (in Units of Δo)
High Spin CFSE (Δo) Low Spin CFSE (Δo)

t2g eg t2g eg

d 0 0

d1 ↿ 0.4

d2 ↿↿ 0.8

d 3 ↿↿↿ 1.2

d4 ↿↿↿ ↿ 0.6 ↿⇂ ↿ ↿ 1.6

d5 ↿↿↿ ↿↿ 0.0 ↿⇂ ↿⇂ ↿ 2.0

d 6 ↿⇂ ↿ ↿ ↿↿ 0.4 ↿⇂ ↿⇂ ↿⇂ 2.4

d7 ↿⇂ ↿⇂ ↿ ↿↿ 0.8 ↿⇂ ↿⇂ ↿⇂ ↿ 1.8

d8 ↿⇂ ↿⇂ ↿⇂ ↿↿ 1.2

d 9 ↿⇂ ↿⇂ ↿⇂ ↿⇂ ↿ 0.6

d 10 ↿⇂ ↿⇂ ↿⇂ ↿⇂ ↿⇂ 0.0

CFSEs are important for two reasons. First, the existence of CFSE nicely accounts for the difference between experimentally
measured values for bond energies in metal complexes and values calculated based solely on electrostatic interactions. Second,
CFSEs represent relatively large amounts of energy (up to several hundred kilojoules per mole), which has important chemical
consequences.

Octahedral d3 and d8 complexes and low-spin d6, d5, d7, and d4 complexes exhibit large
CFSEs.

Summary
Crystal field theory, which assumes that metal–ligand interactions are only electrostatic in nature, explains many important
properties of transition-metal complexes, including their colors, magnetism, structures, stability, and reactivity. Crystal field theory
(CFT) is a bonding model that explains many properties of transition metals that cannot be explained using valence bond theory. In
CFT, complex formation is assumed to be due to electrostatic interactions between a central metal ion and a set of negatively
charged ligands or ligand dipoles arranged around the metal ion. Depending on the arrangement of the ligands, the d orbitals split
into sets of orbitals with different energies. The difference between the energy levels in an octahedral complex is called the crystal
field splitting energy (Δo), whose magnitude depends on the charge on the metal ion, the position of the metal in the periodic table,

3.8.6 https://chem.libretexts.org/@go/page/48469
and the nature of the ligands. The spin-pairing energy (P) is the increase in energy that occurs when an electron is added to an
already occupied orbital. A high-spin configuration occurs when the Δo is less than P, which produces complexes with the
maximum number of unpaired electrons possible. Conversely, a low-spin configuration occurs when the Δo is greater than P, which
produces complexes with the minimum number of unpaired electrons possible. Strong-field ligands interact strongly with the d
orbitals of the metal ions and give a large Δo, whereas weak-field ligands interact more weakly and give a smaller Δo. The colors of
transition-metal complexes depend on the environment of the metal ion and can be explained by CFT.

3.8: Bonding in Octahedral Complex Ions- Crystal Field Theory is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by LibreTexts.

3.8.7 https://chem.libretexts.org/@go/page/48469
3.9: Bonding in Non-Octahedral Complex Ions- Crystal Field Theory
Learning Objectives
Understand the d-orbital degeneracies of square planar and tetrahedral metal complexes.

Tetragonal and Square Planar Complexes


If two trans- ligands in an octahedral complex are either chemically different from the other four (as in trans-[Co(NH3)4Cl2]+), or
at a different distance from the metal than the other four, the result is a tetragonally distorted octahedral complex. The electronic
structures of such complexes are best viewed as the result of distorting an octahedral complex. Consider, for example, an
octahedral complex such as [Co(NH3)6]3+: two trans- NH3 molecules are slowly removed from the metal along the ±z axes, as
shown in the top half of Figure 3.9.1. As the two axial Co–N distances increase simultaneously, the d-orbitals that interact most
strongly with the two axial ligands decrease in energy due to a decrease in electrostatic repulsions between the electrons in these
orbitals and the negative ends of the ligand dipoles. The affected d orbitals are those with a component along the ±z axes—dz2, dxz,
and dyz. These orbitals are not affected equally, however: because the dz2 orbital points directly at the two ligands being removed,
its energy will decrease much more rapidly than the degenerate energies of the dxz and dyz, as shown in the bottom half of Figure
3.9.1. In addition, the effective positive charge on the metal increases somewhat as the axial ligands are removed, increasing the

attraction between the four remaining ligands and the metal. This increases the extent of their interactions with the other two d
orbitals and increases their energies. Again, the two d orbitals are not affected equally: because the dx2−y2 orbital points directly at
the four in-plane ligands, its energy increases to a greater extent than the energy of the dxy orbital, which points between the in-
plane ligands. If the two axial ligands are moved infinitely far away from the metal, a square planar complex is formed. The energy
of the dxy orbital actually surpasses that of the dz2 orbital in the process. The largest orbital splitting in a square planar complex,
between the dx2−y2 and dxy energy levels, is identical in magnitude to Δo.

Figure 3.9.1 d-Orbital Splittings for Tetragonal and Square Planar Complexes. (CC BY-SA-NC; anonymous by request)
Moving the two axial ligands away from the metal ion along the z axis initially generates an elongated octahedral complex (the
center compound of Figure 3.9.1) and eventually produces a square planar complex (right). As shown below the structures, an axial
elongation causes the dz2 dxz and dyz orbitals to decrease in energy and the d 2
x −y
and d orbitals to increase in energy. The
2
xy

change in energy is not the same for all five d orbitals. The dz2 orbital has a smaller xy component than does the dxy orbital, so it
reaches a lower energy level; thus, the order of these orbitals is reversed.

3.9.1 https://chem.libretexts.org/@go/page/81688
Tetrahedral Complexes
In a tetrahedral arrangement of four ligands around a metal ion, none of the ligands lies on any of the three coordinate axes
(illustrated in part (a) in Figure 3.9.2); consequently, none of the five d orbitals points directly at the ligands. Nonetheless, the dxy,
dxz, and dyz orbitals interact more strongly with the ligands than do dx2−y2 and dz2; this again results in a splitting of the five d
orbitals into two groups. The splitting of the energies of the orbitals in a tetrahedral complex (Δt) is much smaller than that for an
octahedral complex (Δo), however, for two reasons: first, the d orbitals interact less strongly with the ligands in a tetrahedral
arrangement; second, there are only four negatively-charged regions rather than six, which decreases the electrostatic interactions
by one-third if all other factors are equal. It can be shown that for complexes of the same metal ion with the same charge, the same
ligands, and the same M–L distance, Δ = Δ . The relationship between the splitting of the five d orbitals in octahedral and
t
4

9
o

tetrahedral crystal fields imposed by the same ligands is shown schematically in part (b) in Figure 3.9.2.

Figure 3.9.2 : d-Orbital Splittings for a Tetrahedral Complex. (a) In a tetrahedral complex, none of the five d orbitals points directly
at or between the ligands. (b) Because the dxy, dxz, and dyzorbitals (the t2g orbitals) interact more strongly with the ligands than do
the dx2−y2 and dz2 orbitals (the eg orbitals), the order of orbital energies in a tetrahedral complex is the opposite of the order in an
octahedral complex. (CC BY-SA-NC; anonymous by request)

Δt < Δo because of weaker d-orbital–ligand interactions and decreased electrostatic


interactions.
Example 3.9.1: Predicting Structure

For each complex, predict its structure, whether it is high spin or low spin, and the number of unpaired electrons present.
1. [C u(N H ) 3 4]
2+

2. [N i(C N ) ]4
2−

Solution
Because Δo is so large for the second- and third-row transition metals, all four-coordinate complexes of these metals are square
planar due to the much higher crystal field stabilization energy (CFSE) for square planar versus tetrahedral structures. The only
exception is for d10 metal ions such as Cd2+, which have zero CFSE and are therefore tetrahedral as predicted by the VSEPR
model. Four-coordinate complexes of the first-row transition metals can be either square planar or tetrahedral; the former is
favored by strong-field ligands, whereas the latter is favored by weak-field ligands. For example, the [Ni(CN)4]2− ion is square
planar, while the [NiCl4]2− ion is tetrahedral.
1.
The copper in this complex is a d9 ion and it has a coordination number of 4. So it is probably either square planar or
tetrahedral. To estimate which, we need to fill in the CFT splitting diagrams for each with 9 electrons and ask which has a
lower energy. Comparing the square planer (Figure 3.9.1) splitting diagram with tetrahedral (Figure 3.9.2), suggests that 9
electrons will have a net lower total energy for square planar (since the d orbital is high in energy, the others are lower).
2
x −y
2

For the square planar structure, neither high nor low spin states are possible (only one state) with a single unpaired electron.
2.
The nickle in this complex is a d8 ion and it has a coordination number of 4. So it is probably either square planar or
tetrahedral. To estimate which, we need to fill in the CFT splitting diagrams for each with 8 electrons and ask which has a

3.9.2 https://chem.libretexts.org/@go/page/81688
lower energy. Comparing the square planer (Figure 3.9.1) splitting diagram with tetrahedral (Figure 3.9.2), suggests that 8
electrons will have a net lower total energy for square planar (since the d orbital is high in energy, the others are lower).
2
x −y
2

For the square planar structure, it is a low spin complex since a high spin requires a lot of energy to promote to the dx2 −y 2

orbital. Hence, there are no unpaired electrons

Exercise 3.9.1

What are the geometries of the following two complexes?


1. [AlCl ] 4

2. [Ag(NH 3
) ]
2
+

Answer 1
tetrahedral
Answer 2
linear

Summary
Distorting an octahedral complex by moving opposite ligands away from the metal produces a tetragonal or square planar
arrangement, in which interactions with equatorial ligands become stronger. Because none of the d orbitals points directly at the
ligands in a tetrahedral complex, these complexes have smaller values of the crystal field splitting energy Δt. In tetrahedral
molecular geometry, a central atom is located at the center of four substituents, which form the corners of a tetrahedron. Tetrahedral
geometry is common for complexes where the metal has d or d electron configuration. The CFT diagram for tetrahedral
0 10

complexes has d and d orbitals equally low in energy because they are between the ligand axis and experience little
x −y
2 2
z
2

repulsion. In square planar molecular geometry, a central atom is surrounded by constituent atoms, which form the corners of a
square on the same plane. The square planar geometry is prevalent for transition metal complexes with d8 configuration. The CFT
diagram for square planar complexes can be derived from octahedral complexes yet the d 2
x −y level is the most destabilized and is
2

left unfilled.

3.9: Bonding in Non-Octahedral Complex Ions- Crystal Field Theory is shared under a CC BY-NC-SA 4.0 license and was authored, remixed,
and/or curated by LibreTexts.
Non-octahedral Complexes by Anonymous is licensed CC BY-NC-SA 4.0.

3.9.3 https://chem.libretexts.org/@go/page/81688
3.10: Magnetic Behavior of Atoms, Molecules, and Materials
Learning Objectives
To understand the difference between Ferromagnetism, paramagnetism and diamagnetism
To identify if a chemical will be paramagnetic or diamagnetic when exposed to an external magnetic field

The magnetic moment of a system measures the strength and the direction of its magnetism. The term itself usually refers to the
magnetic dipole moment. Anything that is magnetic, like a bar magnet or a loop of electric current, has a magnetic moment. A
magnetic moment is a vector quantity, with a magnitude and a direction. An electron has an electron magnetic dipole moment,
generated by the electron's intrinsic spin property, making it an electric charge in motion. There are many different magnetic
behavior including paramagnetism, diamagnetism, and ferromagnetism.
An interesting characteristic of transition metals is their ability to form magnets. Metal complexes that have unpaired electrons are
magnetic. Since the last electrons reside in the d orbitals, this magnetism must be due to having unpaired d electrons. The spin of a
single electron is denoted by the quantum number m as +(1/2) or –(1/2). This spin is negated when the electron is paired with
s

another, but creates a weak magnetic field when the electron is unpaired. More unpaired electrons increase the paramagnetic
effects. The electron configuration of a transition metal (d-block) changes in a coordination compound; this is due to the repulsive
forces between electrons in the ligands and electrons in the compound. Depending on the strength of the ligand, the compound may
be paramagnetic or diamagnetic.

Ferromagnetism (Permanent Magnet)


Ferromagnetism is the basic mechanism by which certain materials (such as iron) form permanent magnets. This means the
compound shows permanent magnetic properties rather than exhibiting them only in the presence of an external magnetic field
(Figure 3.10.1). In a ferromagnetic element, electrons of atoms are grouped into domains in which each domain has the same
charge. In the presence of a magnetic field, these domains line up so that charges are parallel throughout the entire compound.
Whether a compound can be ferromagnetic or not depends on its number of unpaired electrons and on its atomic size.

Figure 3.10.1 : Ferromagnetism (a) nonmagnatized material and (2) Magnetized material with corresponding magnetic fields
shown.

Ferromagnetism, the permanent magnetism associated with nickel, cobalt, and iron, is a common occurrence in everyday life.
Examples of the knowledge and application of ferromagnetism include Aristotle's discussion in 625 BC, the use of the compass
in 1187, and the modern-day refrigerator. Einstein demonstrated that electricity and magnetism are inextricably linked in his
theory of special relativity.

Paramagnetism (Attracted to Magnetic Field)


Paramagnetism refers to the magnetic state of an atom with one or more unpaired electrons. The unpaired electrons are attracted by
a magnetic field due to the electrons' magnetic dipole moments. Hund's Rule states that electrons must occupy every orbital singly
before any orbital is doubly occupied. This may leave the atom with many unpaired electrons. Because unpaired electrons can spin
in either direction, they display magnetic moments in any direction. This capability allows paramagnetic atoms to be attracted to
magnetic fields. Diatomic oxygen, O is a good example of paramagnetism (described via molecular orbital theory). The following
2

video shows liquid oxygen attracted into a magnetic field created by a strong magnet:

3.10.1 https://chem.libretexts.org/@go/page/48470
I'm not a robot
reCAPTCHA
Privacy - Terms

About this page

Our systems have detected unusual traffic from your


computer network. This page checks to see if it's really you
sending the requests, and not a robot. Why did this happen?

IP address: 2604:a880:2:d0::2056:8001
Time: 2023-03-12T20:37:43Z
URL: https://www youtube-
Video 3.10.1 : Paramagnetism of Liquid Oxygen

Diamagnetism (Repelled by Magnetic Field)


As shown in the video, molecular oxygen (O ) is paramagnetic and is attracted to the magnet. In contrast, molecular nitrogen (
2

N ) has no unpaired electrons and is diamagnetic; it is unaffected by the magnet. Diamagnetic substances are characterized by
2

paired electrons, e.g., no unpaired electrons. According to the Pauli Exclusion Principle which states that no two electrons may
occupy the same quantum state at the same time, the electron spins are oriented in opposite directions. This causes the magnetic
fields of the electrons to cancel out; thus there is no net magnetic moment, and the atom cannot be attracted into a magnetic field.
In fact, diamagnetic substances are weakly repelled by a magnetic field as demonstrated with the pyrolytic carbon sheet in Figure
3.10.2.

Figure 3.10.2 : Levitating pyrolytic carbon: A small (~6 mm) piece of pyrolytic graphite levitating over a permanent neodymium
magnet array (5 mm cubes on a piece of steel). Note that the poles of the magnets are aligned vertically and alternate (two with
north facing up, and two with south facing up, diagonally). (Public Domain; Splarka via Wikipedia).

How to Tell if a Substance is Paramagnetic or Diamagnetic


The magnetic properties of a substance can be determined by examining its electron configuration: If it has unpaired electrons,
then the substance is paramagnetic and if all electrons are paired, the substance is then diamagnetic. This process can be broken
into three steps:
1. Write down the electron configuration
2. Draw the valence orbitals
3. Identify if unpaired electrons exist
4. Determine whether the substance is paramagnetic or diamagnetic

3.10.2 https://chem.libretexts.org/@go/page/48470
Example 3.10.1: Chlorine Atoms

Step 1: Find the electron configuration


For Cl atoms, the electron configuration is 3s23p5
Step 2: Draw the valence orbitals
Ignore the core electrons and focus on the valence electrons only.

Step 3: Look for unpaired electrons


There is one unpaired electron.
Step 4: Determine whether the substance is paramagnetic or diamagnetic
Since there is an unpaired electron, Cl atoms are paramagnetic (albeit, weakly).

Example 3.10.2: Zinc Atoms

Step 1: Find the electron configuration


For Zn atoms, the electron configuration is 4s23d10
Step 2: Draw the valence orbitals

Step 3: Look for unpaired electrons


There are no unpaired electrons.
Step 4: Determine whether the substance is paramagnetic or diamagnetic
Because there are no unpaired electrons, Zn atoms are diamagnetic.

Exercise 3.10.1

a. How many unpaired electrons are found in oxygen atoms ?


b. How many unpaired electrons are found in bromine atoms?
c. Indicate whether boron atoms are paramagnetic or diamagnetic.
d. Indicate whether F ions are paramagnetic or diamagnetic.

e. Indicate whether Fe ions are paramagnetic or diamagnetic.


2 +

Answer a
The O atom has 2s22p4 as the electron configuration. Therefore, O has 2 unpaired electrons.

Answer b
The Br atom has 4s23d104p5 as the electron configuration. Therefore, Br has 1 unpaired electron.

Answer c
The B atom has 2s22p1 as the electron configuration. Because it has one unpaired electron, it is paramagnetic.

Answer d
The F ion has 2s22p6 as the electron configuration. Because it has no unpaired electrons, it is diamagnetic.

3.10.3 https://chem.libretexts.org/@go/page/48470
Answer e
The Fe 2 +
ion has 3d6 as the electron configuration. Because it has 4 unpaired electrons, it is paramagnetic.

References
1. Pettrucci, Ralph H. General Chemistry: Principles and Modern Applications. 9th. Upper Saddle River: Pearson Prentice Hall,
2007
2. Sherman, Alan, Sharon J. Sherman, and Leonard Russikoff. Basic Concepts of Chemistry Fifth Edition. Boston, MA: Houghton
Mifflin Company, 1992. Print.

Contributors and Attributions


Jim Clark (Chemguide.co.uk)
Dr. Richard Spinney (The Ohio State University)

3.10: Magnetic Behavior of Atoms, Molecules, and Materials is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by LibreTexts.

3.10.4 https://chem.libretexts.org/@go/page/48470
3.11: Magnetic Behavior of Complex Ions
Learning Objectives
Discuss the correlation between the electronic structure of a coordination complex and its magnetic properties

Experimental evidence of magnetic measurements supports the theory of high- and low-spin complexes. Remember that molecules
such as O that contain unpaired electrons are paramagnetic. Paramagnetic substances are attracted to magnetic fields. Many
2

transition metal complexes have unpaired electrons and hence are paramagnetic. Molecules such as N and ions such as Na and
2
+

[Fe(CN) ]
6
that contain no unpaired electrons are diamagnetic. Diamagnetic substances have a slight tendency to be repelled
4 −

by magnetic fields.
When an electron in an atom or ion is unpaired, the magnetic moment due to its spin makes the entire atom or ion paramagnetic.
The size of the magnetic moment of a system containing unpaired electrons is related directly to the number of such electrons: the
greater the number of unpaired electrons, the larger the magnetic moment. Magnetic susceptibility measures the force experienced
by a substance in a magnetic field. When we compare the weight of a sample to the weight measured in a magnetic field (Figure
3.11.1), paramagnetic samples that are attracted to the magnet will appear heavier because of the force exerted by the magnetic

field. We can calculate the number of unpaired electrons based on the increase in weight.

Figure 3.11.1 : A Gouy balance compares the mass of a sample in the presence of a magnetic field with the mass with the
electromagnet turned off to determine the number of unpaired electrons in a sample. (CC BY-SA 3.0; OpenStax).
From this experiment, the measured magnetic moment of low-spin d6 [Fe(CN) ]4 − ion confirms that iron is diamagnetic,
6

whereas the high-spin d6 [Fe(H2O)6]2+ complex has four unpaired electrons with a magnetic moment that confirms this
arrangement. Therefore, the observed magnetic moment is used to determine the number of unpaired electrons present.

Magnetic Properties of Coordination Compounds


An interesting characteristic of transition metals is their ability to form magnets. Metal complexes that have unpaired electrons are
magnetic. Since the last electrons reside in the d orbitals, this magnetism must be due to having unpaired d electrons. Considering
only monometallic complexes, unpaired electrons arise because the complex has an odd number of electrons or because electron
pairing is destabilized. For example, the Ti(III) ion has one d electron and must be (weakly) paramagnetic, regardless of the
geometry or the nature of the ligands. However, the Ti(II) ion with two d-electrons, sometimes forms complexes with two unpaired
electrons and sometimes forms complexes with no unpaired electrons.
As an example, Fe prefers to exist as Fe and is known to have a coordination number of six. Since the configuration of Fe
3 + 3 +

has five d electrons, we would expect to see five unpaired spins in complexes with Fe. This is true for [FeF ] ; however,
6
3 −

[Fe(CN) ]
6
only has one unpaired electron, making it a weaker magnet. This trend can be explained based on the properties of
3 −

the ligands. Based off the spectrochemical series, we expect CN ligands to have a stronger electric field than that of F ligands,
− −

so the energy differences in the d-orbitals should be greater for the cyanide complex.

3.11.1 https://chem.libretexts.org/@go/page/81695
Figure 3.11.2 : Crystal field theory splitting diagram. Example of influence of ligand electronic properties on d orbital splitting.
This shows the comparison of low-spin versus high-spin electrons.

For this to make sense, there must be some sort of energy benefit to having paired spins for our cyanide complex (the spin pairing
energy). That is, the energy level difference must be more than the repulsive energy of pairing electrons together. Since systems
strive to achieve the lowest energy possible, the electrons will pair up before they will move to the higher orbitals. This is referred
to as low spin, and an electron moving up before pairing is known as high spin.

Exercise 3.11.1

Which ligand is most likely to form a high‐spin complexes when bound to transition metals: en, F‐ , or CN‐ ?

Answer
F

are more likely to form high spin complexes, since is a weak-field ligand on the spectrochemical series.

Tetrahedral complexes have naturally weaker splitting because none of the ligands lie within the plane of the orbitals. As a result,
they have either have too many or too few d electrons to warrant consideration about high or low spin configurations. Square planar
compounds, on the other hand, stem solely from transition metals with eight d electrons. For example, [Ni(CN) ] , 4
2 −

[Pt(NH ) Cl] , and [ PtCl ] are all diamagnetic. Since this encompasses the full spectrum of ligand strength, we can conclude
+ 2 −

3 3 4

that square planar compounds are always low spin and therefore are weakly magnetic.

Example 3.11.2

Which one of the following coordination compounds would you expect to be paramagnetic?
a. [Zn(NH ) ]Cl 3 4 2

b. K[FeCl ] (low spin, tetrahedral)


4

c. [Cd(H O) ]SO
2 6 4

Solution
Let's consider each compound individually. Since each compound has only one complex ion, we only need to consider how the
d electrons are distributed in the ligand field of that ion. We can ignore the other parts (the counter ions) of the compound.
a. The zinc ion in [Zn(NH ) ]Cl has a +2 oxidation state, so it is a d10 ion since the zinc atom loses two 4s electrons and no
3 4 2

3d electrons to form the ion. We do not need to fill ion the electrons in the corresponding tetrahedral d-orbital diagram,
since we know they will all be paired up. This is a diamagnetic compound.
b. The iron ion in K[FeCl ] (low spin, tetrahedral) has a 3+ oxidation state so it is a d5 ion since the iron atom loses two 4s
4

electrons and one 3d electron. Adding 5 electrons to the tetrahedral d-orbital diagram identifies that one electron will be
unpaired. This compound will be paramagnetic (but weakly since only on electron is unpaired).
c. The cadmium ion in [Cd(H O) ]SO has a +2 oxidation state, so it is a d10 ion since the cadmium atom loses two 5s
2 6 4

electrons and no 4d electrons to form the ion. We do not need to fill ion the electrons in the corresponding octahedral d-
orbital diagram, since we know they will all be paired up. This is a diamagnetic compound.

Exercise 3.11.2

Predict the number of unpaired electrons for each of the following complex ions
4 ‐
a. [Fe(CN) ] 6

b. [Ru(NH ) 3 6
3 +
]

3.11.2 https://chem.libretexts.org/@go/page/81695
c. [Cr(NH ) ]
3 6
2 +


d. [EuCl 6
4
]

Answer a
Fe
2 +
is 3d6 ion and has 0 unpaired electron in low spin complexes
Answer b
Ru
3 +
is 4d5 ion and has 1 unpaired electron in low spin complexes
Answer c
Cr
2 +
is 4d4 ion and has 4 and 2 unpaired electrons in high spin and low spin complexes, respectively, depending on spin
pairing energy
Answer d
Eu
2 +
is 4f7 ion and has 7 unpaired electrons in high spin complexes

In bi- and polymetallic complexes, in which the individual centers have an odd number of electrons or electrons are high-spin, the
situation is more complicated. If there is interaction between the two (or more) metal centers, the electrons may couple, resulting in
a weak magnet, or they may enhance each other. When there is no interaction, the two (or more) individual metal centers behave as
if in two separate molecules.

Summary
Unpaired electrons exist when the complex has an odd number of electrons or because electron pairing is destabilized.
The more unpaired electrons, the stronger the magnetic property.
Tetrahedral complexes have weaker splitting because none of the ligands lie within the plane of the orbitals.
Square planar compounds are always low-spin and therefore are weakly magnetic.
In bi- and polymetallic complexes, the electrons may couple through the ligands, resulting in a weak magnet, or they may
enhance each other.

Contributors and Attributions


Source: Boundless. "Magnetic Properties." Boundless Chemistry Boundless, 26 May. 2016. Retrieved 4 May. 2017 from
www.boundless.com/chemistry/...ties-616-6882/.

3.11: Magnetic Behavior of Complex Ions is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

3.11.3 https://chem.libretexts.org/@go/page/81695
3.12: Optical Properties of Coordination Compounds (Color)
The human eye perceives a mixture of all the colors, in the proportions present in sunlight, as white light. Complementary colors, those located
across from each other on a color wheel, are also used in color vision. The eye perceives a mixture of two complementary colors, in the proper
proportions, as white light. Likewise, when a color is missing from white light, the eye sees its complement. For example, when red photons are
absorbed from white light, the eyes see the color green. When violet photons are removed from white light, the eyes see lemon yellow (Figure
3.12.1).

Figure 3.12.1 : (a) An object is black if it absorbs all colors of light. If it reflects all colors of light, it is white. An object has a color if it absorbs all
colors except one, such as this yellow strip. The strip also appears yellow if it absorbs the complementary color from white light (in this case,
indigo). (b) Complementary colors are located directly across from one another on the color wheel. (CC BY 4.0; OpenStax)
The blue color of the [Cu(NH ) ] ion results because this ion absorbs orange and red light, leaving the complementary colors of blue and green
3 4
2 +

(Figure 3.12.2). If white light (ordinary sunlight, for example) passes through [Cu(NH ) ]SO solution, some wavelengths in the light are absorbed
3 4 4
2 +
by the solution. The [Cu(NH ) ] ions in solution absorb light in the red region of the spectrum. The light which passes through the solution and
3 4

out the other side will have all the colors in it except for the red. We see this mixture of wavelengths as pale blue (cyan). The diagram gives an
impression of what happens if you pass white light through a [Cu(NH ) ]SO solution.3 4 4

Figure 3.12.2 : A solution of [Cu(NH3)4]2+ ions absorbs red and orange light, so the transmitted light appears as the complementary color, blue. (CC
BY 4.0; OpenStax)
Working out what color you will see is not easy if you try to do it by imagining "mixing up" the remaining colors. You would not have thought that
all the other colors apart from some red would look cyan, for example. Sometimes what you actually see is quite unexpected. Mixing different
wavelengths of light doesn't give you the same result as mixing paints or other pigments. You can, however, sometimes get some estimate of the
color you would see using the idea of complementary colors.
Recall that the color we observe when we look at an object or a compound is due to light that is transmitted or reflected, not light that is absorbed,
and that reflected or transmitted light is complementary in color to the light that is absorbed. Thus a green compound absorbs light in the red portion
of the visible spectrum and vice versa, as indicated by the complementary color wheel.

Figure 3.12.3 : The partially filled d orbitals of the stable ions Cr3+(aq), Fe3+(aq), and Co2+(aq) (left, center and right, respectively) give rise to
various colors. (credit: Sahar Atwa)
The striking colors exhibited by transition-metal complexes are caused by excitation of an electron from a lower-energy d orbital to a higher-energy
d orbital, which is called a d–d transition (Figure 3.12.4). For a photon to effect such a transition, its energy must be equal to the difference in
energy between the two d orbitals, which depends on the magnitude of Δo.

3.12.1 https://chem.libretexts.org/@go/page/48471
Figure 3.12.4 : In a d–d transition of an octahedral complex, an electron in one of the t2g orbitals of an octahedral complex such as the [Cr(H2O)6]3+
ion absorbs a photon of light with energy equal to Δo, which causes the electron to move to an empty or singly occupied eg orbital.

Example 3.12.1: Colors of Complexes

The octahedral complex [Ti(H2O)6]3+ has a single d electron. To excite this electron from the ground state t2g orbital to the eg orbital, this
complex absorbs light from 450 to 600 nm. The maximum absorbance corresponds to Δo and occurs at 499 nm. Calculate the value of Δo in
Joules and predict what color the solution will appear.
Solution
We can convert wavelength to frequency:
c
ν =
λ

8
3.00 × 10 m/s
=

1m
(499 nm ) × ( )
9
10 nm

14 −1 14
= 6.01 × 10 s = 6.01 × 10 Hz

And then using Planck's equation that related the frequency of light to energy

E = hν

so
−34 14
E = (6.63 × 10 J⋅s) × (6.01 × 10 H z)

−19
= 3.99 × 10 J

Because the complex absorbs 600 nm (orange) through 450 (blue), the indigo, violet, and red wavelengths will be transmitted, and the complex
will appear purple.
Note: This is the energy for one transition (i.e., in one complex). If you want to calculate the energy in J/mol, then you have to multiply this by
Avogadro's number (N ). A

Exercise 3.12.1

A complex that appears green, absorbs photons of what wavelengths?

Answer
red, 620–800 nm

Color Depends on Oxidation State


Small changes in the relative energies of the orbitals that electrons are transitioning between can lead to drastic shifts in the color of light absorbed.
Therefore, the colors of coordination compounds depend on many factors. As shown in Figure 3.12.4, different aqueous metal ions can have
different colors. In addition, different oxidation states of one metal can produce different colors, as shown for the vanadium complexes in the link
below.

3.12.2 https://chem.libretexts.org/@go/page/48471
I'm not a robot
reCAPTCHA
Privacy - Terms

About this page

Our systems have detected unusual traffic from your computer


network. This page checks to see if it's really you sending the
requests, and not a robot. Why did this happen?

IP address: 2604:a880:2:d0::2056:8001
Time: 2023-03-12T20:37:22Z
URL: https://www.youtube-nocookie.com/embed/sFAGQLokym4?
vq=hd1080

Watch this video of the reduction of vanadium complexes to observe the colorful effect of changing oxidation states.

Color Depends on Ligand Field


The specific ligands coordinated to the metal center also influence the color of coordination complexes. Because the energy of a photon of light is
inversely proportional to its wavelength, the color of a complex depends on the magnitude of Δo, which depends on the structure of the complex. For
example, the complex [Cr(NH3)6]3+ has strong-field ligands and a relatively large Δo. Consequently, it absorbs relatively high-energy photons,
corresponding to blue-violet light, which gives it a yellow color. A related complex with weak-field ligands, the [Cr(H2O)6]3+ ion, absorbs lower-
energy photons corresponding to the yellow-green portion of the visible spectrum, giving it a deep violet color. For example, the iron(II) complex
[Fe(H2O)6]SO4 appears blue-green because the high-spin complex absorbs photons in the red wavelengths (Figure 3.12.5). In contrast, the low-spin
iron(II) complex K4[Fe(CN)6] appears pale yellow because it absorbs higher-energy violet photons.

Figure 3.12.5 : Both (a) hexaaquairon(II) sulfate and (b) potassium hexacyanoferrate(II) contain d6 iron(II) octahedral metal centers, but they absorb
photons in different ranges of the visible spectrum. (CC BY 4.0; OpenStax)
In general, strong-field ligands cause a large split in the energies of d orbitals of the central metal atom (large Δoct). Transition metal coordination
compounds with these ligands are yellow, orange, or red because they absorb higher-energy violet or blue light. On the other hand, coordination
compounds of transition metals with weak-field ligands are often blue-green, blue, or indigo because they absorb lower-energy yellow, orange, or
red light.

Strong-field ligands cause a large split in the energies of d orbitals of the central metal atom and
transition metal coordination compounds with these ligands are typically yellow, orange, or red because
they absorb higher-energy violet or blue light. Coordination compounds of transition metals with weak-
field ligands are often blue-green, blue, or indigo because they absorb lower-energy yellow, orange, or
red light.
Example 3.12.2: Matching Colors to Ligand Fields

Identify the color (either blue, green, yellow, or orange) for the following complex ions formed with C o 3+
:
a. [C o(C N ) ] −2
6 3

b. [C o(N H ) ]
3 6
3+

3.12.3 https://chem.libretexts.org/@go/page/48471
c. [C oF ]6
−4

d. [C o(H O) 2 6]
3+

Solution
Each of these complex ions has the same metal with the same oxidation state, so the ligand field is the relevant factors. Each of the complex ions
also has an octahedral ligand field, so we only need to compare the strength of the ligands in determining Δ , which is determined by the o

spectrochemical series.
The ligands for each complex ions are: (a) C N , (b) N H , (c) F

3

and (d) H 2O , which are ranked in increasing Δ magnitude: o

− −
F < H2 O < N H3 < C N (3.12.1)

The relationship between the Δ and the energy of the photons are absorbed in the d-d transition of
o Co
3 +
is given by equating Planck's
equation to the crystal field splitting parameter:
hc
E = hν = = Δo (3.12.2)
λ

Now, we need to get a relative correlation between observed color (to the eye) and the wavelength of the light that is absorbed. From the
complementary color wheel in Figure 3.12.1 we get the following relationships (arranged from highest energy absorbed to lowest):
400-nm Violet light absorbed → Green-yellow observed
430-nm Blue light absorbed → Orange observed
450-nm Blue light absorbed → Yellow observed
490-nm Blue-green light absorbed → Red observed
570-nm Yellow-green light absorbed → Violet observed
580-nm Yellow light absorbed → Dark blue observed
600-nm Orange light absorbed → Blue observed
650-nm Red light absorbed → Green observed
Of the four possible colors given in the problem (blue, green, yellow, and orange), the corresponding colors that are absorbed are (600 nm, 650
nm, 450 nm, and 430, respectively). From Equation 3.12.2, the smaller λ of absorbed light corresponds to the higher energy photons, so we
would correlate the four wavelengths of absorbing photons in terms of increasing energy to observed color:
650 nm < 600 nm < 450 nm < 430 nm (3.12.3)
   
green blue yellow orange

Now, just correlate the ranking in Equation 3.12.3 to the ranking in Equation 3.12.1:
− −
F < H2 O < N H3 < C N
   
green orange
blue yellow

and more specifically in terms of the original question


−4 3+ 3+ −2
[C oF6 ] < [C o(H2 O)6 ] < [C o(N H3 )6 ] < [C o(C N )6 ]
3 3 3
   
green blue yellow orange

Exercise 3.12.1

If a complex ion [M (N H ) ] is red in solution, which of the following ligands, after a ligand exchange reaction to substitute for the ammine
3 6
2+

ligands, may shift the solution to be orange? More than one answer is possible.
Cl

, C N , C O, F
− −
H2 O , I , en , N O , OH , SC N
− −

2
− −

Answer
Shifting from red to orange is an increase in Δ , so those ligands that are stronger field ligands than N H in the list (deterimined via the
o 3

spectrochemical series) are C N , C O, en , and N O .


− −
2

A coordination compound of the Cu+ ion has a d10 configuration, and all the eg orbitals are filled. To excite an electron to a higher level, such as the
4p orbital, photons of very high energy are necessary. This energy corresponds to very short wavelengths in the ultraviolet region of the spectrum.
No visible light is absorbed, so the eye sees no change, and the compound appears white or colorless. A solution containing [Cu(CN)2]−, for
example, is colorless. On the other hand, octahedral Cu2+ complexes have a vacancy in the eg orbitals, and electrons can be excited to this level. The
wavelength (energy) of the light absorbed corresponds to the visible part of the spectrum, and Cu2+ complexes are almost always colored—blue,
blue-green violet, or yellow (Figure 3.12.6). Although CFT successfully describes many properties of coordination complexes, molecular orbital
explanations (beyond the introductory scope provided here) are required to understand fully the behavior of coordination complexes.

3.12.4 https://chem.libretexts.org/@go/page/48471
Figure 3.12.6 : (a) Copper(I) complexes with d10 configurations such as CuI tend to be colorless, whereas (b) d9 copper(II) complexes such as
Cu(NO3)2·5H2O are brightly colored. (CC BY 4.0; OpenStax)

Gems

A ruby is a pink to blood-red colored gemstone that consist of trace amounts of chromium in the mineral corundum Al O . In contrast, 2 3

emeralds are colored green by trace amounts of chromium within a Be3Al2Si6O18 matrix. We can now understand why emeralds and rubies have
such different colors, even though both contain Cr3+ in an octahedral environment provided by six oxide ions. Although the chemical identity of
the six ligands is the same in both cases, the Cr–O distances are different because the compositions of the host lattices are different (Al2O3 in
rubies and Be3Al2Si6O18 in emeralds). In ruby, the Cr–O distances are relatively short because of the constraints of the host lattice, which
increases the d orbital–ligand interactions and makes Δo relatively large. Consequently, rubies absorb green light and the transmitted or reflected
light is red, which gives the gem its characteristic color. In emerald, the Cr–O distances are longer due to relatively large [Si6O18]12− silicate
rings; this results in decreased d orbital–ligand interactions and a smaller Δo. Consequently, emeralds absorb light of a longer wavelength (red),
which gives the gem its characteristic green color. It is clear that the environment of the transition-metal ion, which is determined by the host
lattice, dramatically affects the spectroscopic properties of a metal ion.

Gem-quality crystals of ruby and emerald. The colors of both minerals are due to the presence of small amounts of Cr3+ impurities in octahedral
sites in an otherwise colorless metal oxide lattice.
The absorbance spectrum of a ruby is shown Figure 3.12.7; lef t. The number and positions of the peaks in the spectrum is determined by the
electronic structure of the compound, which in this case depends upon the identity of the metal and the identities, number, and geometry of the
surrounding ions. Crystal field theory may be used to predict the electronic structure and thus the absorbance spectrum. If white light is shown
on the gem, the absorbance spectrum indicates which wavelengths of light are removed. In this case, there are strong bands centered at 414 and
561 nm. These wavelengths correspond with blue and yellow-green light, respectively. For the most part, these colors are not present in the light
reaching ones eyes. An alternative way to express this concept is to recognize that the spectrum of light reaching the eye is the product of the
spectrum of the incident light (white light) and the transmittance spectrum. For this ruby, the transmittance spectrum has a peak at 481 nm and
a broad plateau past 620 nm. (Note that there is significant attenuation of the light across the entire spectrum.) Thus only light with wavelengths
near 481 nm (cyan) and greater than 620 nm (red) reach the eye.

Figure 3.12.7 : (left) Absorbance spectrum of a ruby from Chanthaburi, Thailand. (right) corresponding Transmittance spectrum of a ruby from
Chanthaburi, Thailand. Data obtained from the Caltech Mineral Spectroscopy Server
A similar analysis of the spectrum of an emerald is possible. The absorbance spectrum (Figure 3.12.8; lef t) shows strong bands at 438 and 606
nm, which remove blue and orange light, respectively. The light that is not absorbed is shown by the transmittance spectrum, which indicates the

3.12.5 https://chem.libretexts.org/@go/page/48471
dominant band of light reaching the eye is centered at 512 nm (green light) with smaller contributions from the far blue and red portions of the
spectrum. This combination of wavelengths produces a deep green color.

(left) Absorbance spectrum of an emerald from Malyshevo, Ural, Russia. (right) Corresponding transmittance spectrum of the emerald from
Malyshevo, Ural, Russia. Data obtained from the Caltech Mineral Spectroscopy Server

Summary
When atoms or molecules absorb light at the proper frequency, their electrons are excited to higher-energy orbitals. For many main group atoms and
molecules, the absorbed photons are in the ultraviolet range of the electromagnetic spectrum, which cannot be detected by the human eye. For
coordination compounds, the energy difference between the d orbitals often allows photons in the visible range to be absorbed.

Contributors and Attributions


Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley (Stephen F. Austin
State University) with contributing authors. Textbook content produced by OpenStax College is licensed under a Creative Commons Attribution
License 4.0 license. Download for free at http://cnx.org/contents/[email protected]).
Prof. David Blauch (Davidson College)

3.12: Optical Properties of Coordination Compounds (Color) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
24.7: Color and the Colors of Complexes by OpenStax is licensed CC BY 4.0.

3.12.6 https://chem.libretexts.org/@go/page/48471
CHAPTER OVERVIEW
4: Descriptive Chemistry
UC Davis - Spring 2021

Chemistry 2C
Unit 1: Electrochemistry  ●  Unit 2: Coordination Chemistry  ●  Unit 3: Descriptive Chemistry  ●  Unit 4: Chemical
Kinetics  ●  Unit 5: Nuclear Chemistry
  Agenda  ●  Textbook  ●  ADAPT  ●  Worksheets

Descriptive chemistry addresses the properties of the chemical elements, the compounds that they form, and the underlying
physical and chemical mechanisms involved in their reactivity.
This section is skipped for Spring 2021

4: Descriptive Chemistry is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

4.1 https://chem.libretexts.org/@go/page/13252
CHAPTER OVERVIEW
5: Chemical Kinetics
UC Davis - Spring 2021

Chemistry 2C
Unit 1: Electrochemistry  ●  Unit 2: Coordination Chemistry  ●  Unit 3: Descriptive Chemistry  ●  Unit 4: Chemical
Kinetics  ●  Unit 5: Nuclear Chemistry
  Agenda  ●  Textbook  ●  ADAPT  ●  Worksheets

This chapter will present a quantitative description of when the chemical composition of a system is not constant with time.
Chemical kinetics is the study of reaction rates, the changes in the concentrations of reactants and products with time. With a
discussion of chemical kinetics, the reaction rates or the changes in the concentrations of reactants and products with time are
studied. The techniques you are about to learn will enable you to describe the speed of many such changes and predict how the
composition of each system will change in response to changing conditions. As you learn about the factors that affect reaction
rates, the methods chemists use for reporting and calculating those rates, and the clues that reaction rates provide about events at
the molecular level.
5.1: Factors that Affect Reaction Rates
5.2: Reaction Rates
5.3: Concentration and Rates (Rate Laws)
5.4: Determining Rate Laws from Initial Rates (Differential Rate Laws)
5.5: The Change of Concentration with Time (Integrated Rate Laws)
5.6: Using Graphs to Determine (Integrated) Rate Laws
5.7: Collision Theory
5.8: Temperature and Rate
5.9: Reaction Mechanisms
5.10: Catalysis
5.11: Exercises
5.12: Chemical Kinetics (Summary)

5: Chemical Kinetics is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

1
5.1: Factors that Affect Reaction Rates
 Learning Objectives
To gain a general overview of some of the important factors that manipulate chemical reaction rates.

There are many factors that influence the reaction rates of chemical reactions include the concentration of reactants, temperature,
the physical state of reactants and their dispersion, the solvent, and the presence of a catalyst.
Although a balanced chemical equation for a reaction describes the quantitative relationships between the amounts of reactants
present and the amounts of products that can be formed, it gives us no information about whether or how fast a given reaction will
occur. This information is obtained by studying the chemical kinetics of a reaction, which depend on various factors: reactant
concentrations, temperature, physical states and surface areas of reactants, and solvent and catalyst properties if either are present.
By studying the kinetics of a reaction, chemists gain insights into how to control reaction conditions to achieve a desired outcome.

Concentration Effects
Two substances cannot possibly react with each other unless their constituent particles (molecules, atoms, or ions) come into
contact. If there is no contact, the reaction rate will be zero. Conversely, the more reactant particles that collide per unit time, the
more often a reaction between them can occur. Consequently, the reaction rate usually increases as the concentration of the
reactants increases.

Temperature Effects
Increasing the temperature of a system increases the average kinetic energy of its constituent particles. As the average kinetic
energy increases, the particles move faster and collide more frequently per unit time and possess greater energy when they collide.
Both of these factors increase the reaction rate. Hence the reaction rate of virtually all reactions increases with increasing
temperature. Conversely, the reaction rate of virtually all reactions decreases with decreasing temperature. For example,
refrigeration retards the rate of growth of bacteria in foods by decreasing the reaction rates of biochemical reactions that enable
bacteria to reproduce.
In systems where more than one reaction is possible, the same reactants can produce different products under different reaction
conditions. For example, in the presence of dilute sulfuric acid and at temperatures around 100°C, ethanol is converted to diethyl
ether:
H2 SO4

2C H3 C H2 OH −−−−→ C H3 C H2 OC H2 C H3 + H2 O (5.1.1)

At 180°C, however, a completely different reaction occurs, which produces ethylene as the major product:
H2 SO4

C H3 C H2 OH −−−−→ C2 H4 + H2 O (5.1.2)

Phase and Surface Area Effects


When two reactants are in the same fluid phase, their particles collide more frequently than when one or both reactants are solids
(or when they are in different fluids that do not mix). If the reactants are uniformly dispersed in a single homogeneous solution,
then the number of collisions per unit time depends on concentration and temperature, as we have just seen. If the reaction is
heterogeneous, however, the reactants are in two different phases, and collisions between the reactants can occur only at interfaces
between phases. The number of collisions between reactants per unit time is substantially reduced relative to the homogeneous
case, and, hence, so is the reaction rate. The reaction rate of a heterogeneous reaction depends on the surface area of the more
condensed phase.
Automobile engines use surface area effects to increase reaction rates. Gasoline is injected into each cylinder, where it combusts on
ignition by a spark from the spark plug. The gasoline is injected in the form of microscopic droplets because in that form it has a
much larger surface area and can burn much more rapidly than if it were fed into the cylinder as a stream. Similarly, a pile of finely
divided flour burns slowly (or not at all), but spraying finely divided flour into a flame produces a vigorous reaction.

5.1.1 https://chem.libretexts.org/@go/page/81880
Solvent Effects
The nature of the solvent can also affect the reaction rates of solute particles. For example, a sodium acetate solution reacts with
methyl iodide in an exchange reaction to give methyl acetate and sodium iodide.

C H3 C O2 N a(soln) + C H3 I(l) → C H3 C O2 C H3 (soln) + N aI(soln) (5.1.3)

This reaction occurs 10 million times more rapidly in the organic solvent dimethylformamide [DMF; (CH3)2NCHO] than it does in
methanol (CH3OH). Although both are organic solvents with similar dielectric constants (36.7 for DMF versus 32.6 for methanol),
methanol is able to hydrogen bond with acetate ions, whereas DMF cannot. Hydrogen bonding reduces the reactivity of the oxygen
atoms in the acetate ion.
Solvent viscosity is also important in determining reaction rates. In highly viscous solvents, dissolved particles diffuse much more
slowly than in less viscous solvents and can collide less frequently per unit time. Thus the reaction rates of most reactions decrease
rapidly with increasing solvent viscosity.

Catalyst Effects
A catalyst is a substance that participates in a chemical reaction and increases the reaction rate without undergoing a net chemical
change itself. Consider, for example, the decomposition of hydrogen peroxide in the presence and absence of different catalysts.
Because most catalysts are highly selective, they often determine the product of a reaction by accelerating only one of several
possible reactions that could occur.
Most of the bulk chemicals produced in industry are formed with catalyzed reactions. Recent estimates indicate that about 30% of
the gross national product of the United States and other industrialized nations relies either directly or indirectly on the use of
catalysts.

5.1: Factors that Affect Reaction Rates is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
14.1: Factors that Affect Reaction Rates is licensed CC BY-NC-SA 3.0.

5.1.2 https://chem.libretexts.org/@go/page/81880
5.2: Reaction Rates
 Learning Objectives
To determine the reaction rate of a reaction.

Reaction rates are usually expressed as the concentration of reactant consumed or the concentration of product formed per unit
time. The units are thus moles per liter per unit time, written as M/s, M/min, or M/h. To measure reaction rates, chemists initiate the
reaction, measure the concentration of the reactant or product at different times as the reaction progresses, perhaps plot the
concentration as a function of time on a graph, and then calculate the change in the concentration per unit time.

Figure 5.2.1 : The Progress of a Simple Reaction (A → B). The mixture initially contains only A molecules (purple). Over time, the
number of A molecules decreases and more B molecules (green) are formed (top). The graph shows the change in the number of A
and B molecules in the reaction as a function of time over a 1 min period (bottom).
The progress of a simple reaction (A → B) is shown in Figure 5.2.1; the beakers are snapshots of the composition of the solution at
10 s intervals. The number of molecules of reactant (A) and product (B) are plotted as a function of time in the graph. Each point in
the graph corresponds to one beaker in Figure 5.2.1. The reaction rate is the change in the concentration of either the reactant or the
product over a period of time. The concentration of A decreases with time, while the concentration of B increases with time.
Δ[B] Δ[A]
rate = =− (5.2.1)
Δt Δt

Square brackets indicate molar concentrations, and the capital Greek delta (Δ) means “change in.” Because chemists follow the
convention of expressing all reaction rates as positive numbers, however, a negative sign is inserted in front of Δ[A]/Δt to convert
that expression to a positive number. The reaction rate calculated for the reaction A → B using Equation 5.2.1 is different for each
interval (this is not true for every reaction, as shown below). A greater change occurs in [A] and [B] during the first 10 s interval,
for example, than during the last, meaning that the reaction rate is greatest at first.

Reaction rates generally decrease with time as reactant concentrations decrease.

5.2.1 https://chem.libretexts.org/@go/page/81881
I'm not a robot
reCAPTCHA
Privacy - Terms

About this page

Our systems have detected unusual traffic from your


computer network This page checks to see if it's really you
A Video Discussing Average Reaction Rates. Video Link: Introduction to Chemical Reaction Kinetics(opens in new window)
[youtu.be] (opens in new window)

Determining the Reaction Rate of Hydrolysis of Aspirin


We can use Equation 5.2.1 to determine the reaction rate of hydrolysis of aspirin, probably the most commonly used drug in the
world (more than 25,000,000 kg are produced annually worldwide). Aspirin (acetylsalicylic acid) reacts with water (such as water
in body fluids) to give salicylic acid and acetic acid, as shown in Figure 5.2.2.

Figure 5.2.2 : Hydrolysis of Aspirin reaction.


Because salicylic acid is the actual substance that relieves pain and reduces fever and inflammation, a great deal of research has
focused on understanding this reaction and the factors that affect its rate. Data for the hydrolysis of a sample of aspirin are in Table
5.2.1 and are shown in the graph in Figure 5.2.3.

Table 5.2.1 : Data for Aspirin Hydrolysis in Aqueous Solution at pH 7.0 and 37°C*
Time (h) [Aspirin] (M) [Salicylic Acid] (M)

0 5.55 × 10−3 0

2.0 5.51 × 10−3 0.040 × 10−3

5.0 5.45 × 10−3 0.10 × 10−3

10 5.35 × 10−3 0.20 × 10−3

20 5.15 × 10−3 0.40 × 10−3

30 4.96 × 10−3 0.59 × 10−3

40 4.78 × 10−3 0.77 × 10−3

50 4.61 × 10−3 0.94 × 10−3

100 3.83 × 10−3 1.72 × 10−3

200 2.64 × 10−3 2.91 × 10−3

300 1.82 × 10−3 3.73 × 10−3

*The reaction at pH 7.0 is very slow. It is much faster under acidic conditions, such as those found in the stomach.

The data in Table 5.2.1 were obtained by removing samples of the reaction mixture at the indicated times and analyzing them for
the concentrations of the reactant (aspirin) and one of the products (salicylic acid).

5.2.2 https://chem.libretexts.org/@go/page/81881
Figure 5.2.3 : The Hydrolysis of Aspirin. This graph shows the concentrations of aspirin and salicylic acid as a function of time,
based on the hydrolysis data in Table 14.1. The time dependence of the concentration of the other product, acetate, is not shown,
but based on the stoichiometry of the reaction, it is identical to the data for salicylic acid.

Graph of concentration against time in hours. The purple line is aspirin. The green line is salicylic acid.

The average reaction rate for a given time interval can be calculated from the concentrations of either the reactant or one of the
products at the beginning of the interval (time = t0) and at the end of the interval (t1). Using salicylic acid, the reaction rate for the
interval between t = 0 h and t = 2.0 h (recall that change is always calculated as final minus initial) is calculated as follows:
[salicyclic acid]2 − [salicyclic acid]0
rate(t=0−2.0 h) =
2.0 h − 0 h
−3
0.040 × 10  M − 0 M −5
= = 2.0 × 10  M/h
2.0 h

The reaction rate can also be calculated from the concentrations of aspirin at the beginning and the end of the same interval,
remembering to insert a negative sign, because its concentration decreases:
[aspirin]2 − [aspirin]0
rate(t=0−2.0 h) = −
2.0 h − 0 h
−3 −3
(5.51 × 10  M) − (5.55 × 10  M)
=−
2.0 h
−5
= 2 × 10  M/h

If the reaction rate is calculated during the last interval given in Table 5.2.1(the interval between 200 h and 300 h after the start of
the reaction), the reaction rate is significantly slower than it was during the first interval (t = 0–2.0 h):
[salicyclic acid]300 − [salicyclic acid]200
rate(t=200−300h) =
300 h − 200 h
−3 −3
(3.73 × 10  M) − (2.91 × 10  M)
=−
100 h
−6
= 8.2 × 10  M/h

Calculating the Reaction Rate of Fermentation of Sucrose


In the preceding example, the stoichiometric coefficients in the balanced chemical equation are the same for all reactants and
products; that is, the reactants and products all have the coefficient 1. Consider a reaction in which the coefficients are not all the
same, the fermentation of sucrose to ethanol and carbon dioxide:

5.2.3 https://chem.libretexts.org/@go/page/81881
C12 H22 O11 (aq) + H2 O(l) → 4 C2 H5 OH(aq) + 4C O2 (g) (5.2.2)
sucrose

The coefficients indicate that the reaction produces four molecules of ethanol and four molecules of carbon dioxide for every one
molecule of sucrose consumed. As before, the reaction rate can be found from the change in the concentration of any reactant or
product. In this particular case, however, a chemist would probably use the concentration of either sucrose or ethanol because gases
are usually measured as volumes and, as explained in Chapter 10, the volume of CO2 gas formed depends on the total volume of
the solution being studied and the solubility of the gas in the solution, not just the concentration of sucrose. The coefficients in the
balanced chemical equation tell us that the reaction rate at which ethanol is formed is always four times faster than the reaction rate
at which sucrose is consumed:
Δ[ C2 H5 OH] 4Δ[sucrose]
=− (5.2.3)
Δt Δt

The concentration of the reactant—in this case sucrose—decreases with time, so the value of Δ[sucrose] is negative. Consequently,
a minus sign is inserted in front of Δ[sucrose] in Equation 5.2.3 so the rate of change of the sucrose concentration is expressed as a
positive value. Conversely, the ethanol concentration increases with time, so its rate of change is automatically expressed as a
positive value.
Often the reaction rate is expressed in terms of the reactant or product with the smallest coefficient in the balanced chemical
equation. The smallest coefficient in the sucrose fermentation reaction (Equation 5.2.2) corresponds to sucrose, so the reaction rate
is generally defined as follows:
Δ[sucrose] 1 Δ[ C2 H5 OH]
rate = − = ( ) (5.2.4)
Δt 4 Δt

 Example 5.2.1: Decomposition Reaction I

Consider the thermal decomposition of gaseous N2O5 to NO2 and O2 via the following equation:
Δ

2 N2 O5 (g) −→ 4NO2 (g) + O2 (g)

Write expressions for the reaction rate in terms of the rates of change in the concentrations of the reactant and each product
with time.
Given: balanced chemical equation
Asked for: reaction rate expressions

Strategy:
A. Choose the species in the equation that has the smallest coefficient. Then write an expression for the rate of change of that
species with time.
B. For the remaining species in the equation, use molar ratios to obtain equivalent expressions for the reaction rate.

Solution
A Because O2 has the smallest coefficient in the balanced chemical equation for the reaction, define the reaction rate as the rate
of change in the concentration of O2 and write that expression.
B The balanced chemical equation shows that 2 mol of N2O5 must decompose for each 1 mol of O2 produced and that 4 mol of
NO2 are produced for every 1 mol of O2 produced. The molar ratios of O2 to N2O5 and to NO2 are thus 1:2 and 1:4,
respectively. This means that the rate of change of [N2O5] and [NO2] must be divided by its stoichiometric coefficient to obtain
equivalent expressions for the reaction rate. For example, because NO2 is produced at four times the rate of O2, the rate of
production of NO2 is divided by 4. The reaction rate expressions are as follows:
Δ[ O2 ] Δ[NO2 ] Δ[ N2 O5 ]
rate = = =−
Δt 4Δt 2Δt

5.2.4 https://chem.libretexts.org/@go/page/81881
 Exercise 5.2.1: Contact Process I

The contact process is used in the manufacture of sulfuric acid. A key step in this process is the reaction of SO2 with O to
2

produce SO . 3

2S O2(g) + O2(g) → 2S O3(g)

Write expressions for the reaction rate in terms of the rate of change of the concentration of each species.

Answer
Δ[ O2 ] Δ[SO2 ] Δ[SO3 ]
rate = − =− =
Δt 2Δt 2Δt

Instantaneous Rates of Reaction


The instantaneous rate of a reaction is the reaction rate at any given point in time. As the period of time used to calculate an
average rate of a reaction becomes shorter and shorter, the average rate approaches the instantaneous rate. Comparing this to
calculus, the instantaneous rate of a reaction at a given time corresponds to the slope of a line tangent to the concentration-versus-
time curve at that point—that is, the derivative of concentration with respect to time.
The distinction between the instantaneous and average rates of a reaction is similar to the distinction between the actual speed of a
car at any given time on a trip and the average speed of the car for the entire trip. Although the car may travel for an extended
period at 65 mph on an interstate highway during a long trip, there may be times when it travels only 25 mph in construction zones
or 0 mph if you stop for meals or gas. The average speed on the trip may be only 50 mph, whereas the instantaneous speed on the
interstate at a given moment may be 65 mph. Whether the car can be stopped in time to avoid an accident depends on its
instantaneous speed, not its average speed. There are important differences between the speed of a car during a trip and the speed of
a chemical reaction, however. The speed of a car may vary unpredictably over the length of a trip, and the initial part of a trip is
often one of the slowest. In a chemical reaction, the initial interval typically has the fastest rate (though this is not always the case),
and the reaction rate generally changes smoothly over time.

Chemical kinetics generally focuses on one particular instantaneous rate, which is the
initial reaction rate, t = 0. Initial rates are determined by measuring the reaction rate at
various times and then extrapolating a plot of rate versus time to t = 0.

I'm not a robot


reCAPTCHA
Privacy - Terms

About this page

Our systems have detected unusual traffic from your


computer network This page checks to see if it's really you
 Example 5.2.2: Decomposition Reaction II

Using the reaction shown in Example 5.2.1, calculate the reaction rate from the following data taken at 56°C:

2 N2 O5(g) → 4N O2(g) + O2(g)

calculate the reaction rate from the following data taken at 56°C:
Time (s) [N2O5] (M) [NO2] (M) [O2] (M)

5.2.5 https://chem.libretexts.org/@go/page/81881
Time (s) [N2O5] (M) [NO2] (M) [O2] (M)

240 0.0388 0.0314 0.00792

600 0.0197 0.0699 0.0175

Given: balanced chemical equation and concentrations at specific times


Asked for: reaction rate

Strategy:
A. Using the equations in Example 5.2.1, subtract the initial concentration of a species from its final concentration and
substitute that value into the equation for that species.
B. Substitute the value for the time interval into the equation. Make sure your units are consistent.

Solution
A Calculate the reaction rate in the interval between t1 = 240 s and t2 = 600 s. From Example 5.2.1 , the reaction rate can be
evaluated using any of three expressions:
Δ[ O2 ] Δ[NO2 ] Δ[ N2 O5 ]
rate = = =−
Δt 4Δt 2Δt

Subtracting the initial concentration from the final concentration of N2O5 and inserting the corresponding time interval into the
rate expression for N2O5,
Δ[ N2 O5 ] [ N2 O5 ]600 − [ N2 O5 ]240
rate = − =−
2Δt 2(600 s − 240 s)

B Substituting actual values into the expression,


0.0197 M − 0.0388 M
−5
rate = − = 2.65 × 10  M/s
2(360 s)

Similarly, NO2 can be used to calculate the reaction rate:


Δ[NO2 ] [NO2 ]600 − [NO2 ]240 0.0699 M − 0.0314 M −5
rate = = = = 2.67 × 10 M/s
4Δt 4(600 s − 240 s) 4(360 s)

Allowing for experimental error, this is the same rate obtained using the data for N2O5. The data for O2 can also be used:
Δ[ O2 ] [ O2 ]600 − [ O2 ]240 0.0175 M − 0.00792 M −5
rate = = = = 2.66 × 10 M/s
Δt 600 s − 240 s 360 s

Again, this is the same value obtained from the N2O5 and NO2 data. Thus, the reaction rate does not depend on which reactant
or product is used to measure it.

 Exercise 5.2.2: Contact Process II


Using the data in the following table, calculate the reaction rate of SO 2 (g) with O 2 (g) to give SO 3 (g) .

2S O2(g) + O2(g) → 2S O3(g)

calculate the reaction rate of SO 2 (g) with O2 (g) to give SO .


3 (g)

Time (s) [SO2] (M) [O2] (M) [SO3] (M)

300 0.0270 0.0500 0.0072

720 0.0194 0.0462 0.0148

5.2.6 https://chem.libretexts.org/@go/page/81881
Answer:
9.0 × 10−6 M/s

Summary
In this Module, the quantitative determination of a reaction rate is demonstrated. Reaction rates can be determined over particular
time intervals or at a given point in time. A rate law describes the relationship between reactant rates and reactant concentrations.
Reaction rates are reported as either the average rate over a period of time or as the instantaneous rate at a single time. Reaction
rates can be determined over particular time intervals or at a given point in time.
General definition of rate for A → B:
Δ[B] Δ[A]
rate = =−
Δt Δt

5.2: Reaction Rates is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
14.2: Reaction Rates is licensed CC BY-NC-SA 3.0.

5.2.7 https://chem.libretexts.org/@go/page/81881
5.3: Concentration and Rates (Rate Laws)
 Learning Objectives
To understand the meaning of the rate law.

The factors that affect the reaction rate of a chemical reaction, which may determine whether a desired product is formed. In this
section, we will show you how to quantitatively determine the reaction rate.

Rate Laws
Typically, reaction rates decrease with time because reactant concentrations decrease as reactants are converted to products.
Reaction rates generally increase when reactant concentrations are increased. This section examines mathematical expressions
called rate laws, which describe the relationships between reactant rates and reactant concentrations. Rate laws are mathematical
descriptions of experimentally verifiable data.
Rate laws may be written from either of two different but related perspectives. A differential rate law expresses the reaction rate
in terms of changes in the concentration of one or more reactants (Δ[R]) over a specific time interval (Δt). In contrast, an
integrated rate law describes the reaction rate in terms of the initial concentration ([R]0) and the measured concentration of one or
more reactants ([R]) after a given amount of time (t); integrated rate laws are discussed in more detail later. The integrated rate law
is derived by using calculus to integrate the differential rate law. Whether using a differential rate law or integrated rate law, always
make sure that the rate law gives the proper units for the reaction rate, usually moles per liter per second (M/s).

Reaction Orders
For a reaction with the general equation:
aA + bB → cC + dD (5.3.1)

the experimentally determined rate law usually has the following form:
m n
rate = k[A] [B] (5.3.2)

The proportionality constant (k) is called the rate constant, and its value is characteristic of the reaction and the reaction
conditions. A given reaction has a particular rate constant value under a given set of conditions, such as temperature, pressure, and
solvent; varying the temperature or the solvent usually changes the value of the rate constant. The numerical value of k, however,
does not change as the reaction progresses under a given set of conditions.

Under a given set of conditions, the value of the rate constant does not change as the
reaction progresses.
The reaction rate thus depends on the rate constant for the given set of reaction conditions and the concentration of A and B raised
to the powers m and n, respectively. The values of m and n are derived from experimental measurements of the changes in reactant
concentrations over time and indicate the reaction order, the degree to which the reaction rate depends on the concentration of
each reactant; m and n need not be integers. For example, Equation 5.3.2 tells us that Equation 5.3.1 is mth order in reactant A and
nth order in reactant B. It is important to remember that n and m are not related to the stoichiometric coefficients a and b in the
balanced chemical equation and must be determined experimentally. The overall reaction order is the sum of all the exponents in
the rate law: m + n.

The orders of the reactions (e.g. n and m) are not related to the stoichiometric coefficients
in the balanced chemical (e.g., a and b).
To illustrate how chemists interpret a differential rate law, consider the experimentally derived rate law for the hydrolysis of t-butyl
bromide in 70% aqueous acetone.

5.3.1 https://chem.libretexts.org/@go/page/81882
This reaction produces t-butanol according to the following equation:
(C H3 )3 C Br(soln) + H2 O(soln) → (C H3 )3 C OH(soln) + H Br(soln) (5.3.3)

Combining the rate expression in Equation 5.3.2 with the definition of average reaction rate
Δ[A]
rate = −
Δt

gives a general expression for the differential rate law:


Δ[A]
m n
rate = − = k[A] [B] (5.3.4)
Δt

Inserting the identities of the reactants into Equation 5.3.4 gives the following expression for the differential rate law for the
reaction:
Δ[(C H3 )3 CBr]
m n
rate = − = k[(C H3 )3 CBr] [ H2 O] (5.3.5)
Δt

Experiments to determine the rate law for the hydrolysis of t-butyl bromide show that the reaction rate is directly proportional to
the concentration of (CH3)3CBr but is independent of the concentration of water. Therefore, m and n in Equation 5.3.4 are 1 and 0,
respectively, and,
1 0
rate = k[(C H3 )3 C Br] [ H2 O] = k[(C H3 )3 C Br] (5.3.6)

Because the exponent for the reactant is 1, the reaction is first order in (CH3)3CBr. It is zeroth order in water because the exponent
for [H2O] is 0. (Recall that anything raised to the zeroth power equals 1.) Thus, the overall reaction order is 1 + 0 = 1. The reaction
orders state in practical terms that doubling the concentration of (CH3)3CBr doubles the reaction rate of the hydrolysis reaction,
halving the concentration of (CH3)3CBr halves the reaction rate, and so on. Conversely, increasing or decreasing the concentration
of water has no effect on the reaction rate. (Again, when working with rate laws, there is no simple correlation between the
stoichiometry of the reaction and the rate law. The values of k, m, and n in the rate law must be determined experimentally.)
Experimental data show that k has the value 5.15 × 10−4 s−1 at 25°C. The rate constant has units of reciprocal seconds (s−1) because
the reaction rate is defined in units of concentration per unit time (M/s). The units of a rate constant depend on the rate law for a
particular reaction.
Under conditions identical to those for the t-butyl bromide reaction, the experimentally derived differential rate law for the
hydrolysis of methyl bromide (CH3Br) is as follows:
Δ[C H3 Br]

rate = − = k [C H3 Br] (5.3.7)
Δt

This reaction also has an overall reaction order of 1, but the rate constant in Equation 5.3.7 is approximately 106 times smaller than
that for t-butyl bromide. Thus, methyl bromide hydrolyzes about 1 million times more slowly than t-butyl bromide, and this
information tells chemists how the reactions differ on a molecular level.
Frequently, changes in reaction conditions also produce changes in a rate law. In fact, chemists often alter reaction conditions to
study the mechanics of a reaction. For example, when t-butyl bromide is hydrolyzed in an aqueous acetone solution containing
OH− ions rather than in aqueous acetone alone, the differential rate law for the hydrolysis reaction does not change. In contrast, for
methyl bromide, the differential rate law becomes

rate = k'' [C H3 Br][OH ]

with an overall reaction order of 2. Although the two reactions proceed similarly in neutral solution, they proceed very differently
in the presence of a base, providing clues as to how the reactions differ on a molecular level.

5.3.2 https://chem.libretexts.org/@go/page/81882
 Example 5.3.1: Writing Rate Laws from Reaction Orders

An experiment shows that the reaction of nitrogen dioxide with carbon monoxide:
NO (g) + CO(g) ⟶ NO(g) + CO (g)
2 2

is second order in NO2 and zero order in CO at 100 °C. What is the rate law for the reaction?

Solution
The reaction will have the form:
m n
rate = k[ NO ] [CO]
2

The reaction is second order in NO2; thus m = 2. The reaction is zero order in CO; thus n = 0. The rate law is:
2 0 2
rate = k[ NO ] [CO] = k[ NO ]
2 2

Remember that a number raised to the zero power is equal to 1, thus [CO]0 = 1, which is why we can simply drop the
concentration of CO from the rate equation: the rate of reaction is solely dependent on the concentration of NO2. When we
consider rate mechanisms later in this chapter, we will explain how a reactant’s concentration can have no effect on a reaction
despite being involved in the reaction.

 Exercise 5.3.1A

The rate law for the reaction:

H (g) + 2 NO(g) ⟶ N O(g) + H O(g)


2 2 2

has been experimentally determined to be rate = k[N O] 2


[ H2 ] . What are the orders with respect to each reactant, and what is the
overall order of the reaction?

Answer
order in NO = 2
order in H2 = 1
overall order = 3

 Exercise 5.3.1B

In a transesterification reaction, a triglyceride reacts with an alcohol to form an ester and glycerol. Many students learn about
the reaction between methanol (CH3OH) and ethyl acetate (CH3CH2OCOCH3) as a sample reaction before studying the
chemical reactions that produce biodiesel:
CH OH + CH CH OCOCH ⟶ CH OCOCH + CH CH OH
3 3 2 3 3 3 3 2

The rate law for the reaction between methanol and ethyl acetate is, under certain conditions, experimentally determined to be:

rate = k[ CH OH]
3

What is the order of reaction with respect to methanol and ethyl acetate, and what is the overall order of reaction?

Answer
order in CH3OH = 1
order in CH3CH2OCOCH3 = 0
overall order = 1

5.3.3 https://chem.libretexts.org/@go/page/81882
 Example 5.3.2: Differential Rate Laws
Below are three reactions and their experimentally determined differential rate laws. For each reaction, give the units of the
rate constant, give the reaction order with respect to each reactant, give the overall reaction order, and predict what happens to
the reaction rate when the concentration of the first species in each chemical equation is doubled.
Pt

a. 2HI(g) −→ H2 (g) + I2 (g)

1 Δ[HI]
2
rate = − ( ) = k[HI]
2 Δt

b. 2 N2 O(g) −
→ 2 N2 (g) + O2 (g)

1 Δ[ N2 O]
rate = − ( ) =k
2 Δt

c. cyclopropane(g) → propane(g)

Δ[cyclopropane]
rate = − = k[cyclopropane]
Δt

Given: balanced chemical equations and differential rate laws


Asked for: units of rate constant, reaction orders, and effect of doubling reactant concentration

Strategy:
A. Express the reaction rate as moles per liter per second [mol/(L·s), or M/s]. Then determine the units of each chemical
species in the rate law. Divide the units for the reaction rate by the units for all species in the rate law to obtain the units for
the rate constant.
B. Identify the exponent of each species in the rate law to determine the reaction order with respect to that species. Add all
exponents to obtain the overall reaction order.
C. Use the mathematical relationships as expressed in the rate law to determine the effect of doubling the concentration of a
single species on the reaction rate.

Solution
1. A [HI]2 will give units of (moles per liter)2. For the reaction rate to have units of moles per liter per second, the rate
constant must have reciprocal units [1/(M·s)]:
M M/s 1
2 −1 −1
kM = k = = =M ⋅s
2
s M M⋅s

B The exponent in the rate law is 2, so the reaction is second order in HI. Because HI is the only reactant and the only
species that appears in the rate law, the reaction is also second order overall.
C If the concentration of HI is doubled, the reaction rate will increase from k[HI]02 to k(2[HI])02 = 4k[HI]02. The reaction
rate will therefore quadruple.
2. A Because no concentration term appears in the rate law, the rate constant must have M/s units for the reaction rate to have
M/s units.
B The rate law tells us that the reaction rate is constant and independent of the N2O concentration. That is, the reaction is
zeroth order in N2O and zeroth order overall.
C Because the reaction rate is independent of the N2O concentration, doubling the concentration will have no effect on the
reaction rate.
3. A The rate law contains only one concentration term raised to the first power. Hence the rate constant must have units of
reciprocal seconds (s−1) to have units of moles per liter per second for the reaction rate: M·s−1 = M/s.

5.3.4 https://chem.libretexts.org/@go/page/81882
B The only concentration in the rate law is that of cyclopropane, and its exponent is 1. This means that the reaction is first
order in cyclopropane. Cyclopropane is the only species that appears in the rate law, so the reaction is also first order
overall.
C Doubling the initial cyclopropane concentration will increase the reaction rate from k[cyclopropane]0 to
2k[cyclopropane]0. This doubles the reaction rate.

 Exercise 5.3.2

Given the following two reactions and their experimentally determined differential rate laws: determine the units of the rate
constant if time is in seconds, determine the reaction order with respect to each reactant, give the overall reaction order, and
predict what will happen to the reaction rate when the concentration of the first species in each equation is doubled.
Δ[ CH3 N=NCH3 ]
CH3 N=NCH3 (g) → C2 H6 (g) + N2 (g) rate = − (5.3.8)
a. Δt

= k[ CH3 N=NCH3 ] (5.3.9)

Δ[ F2 ] 1 Δ[NO2 ]
2NO2 (g) + F2 (g) → 2NO2 F(g) rate = − =− ( ) (5.3.10)
b. Δt 2 Δt

= k[NO2 ][ F2 ] (5.3.11)

Answer a
s−1; first order in CH3N=NCH3; first order overall; doubling [CH3N=NCH3] will double the reaction rate.
Answer b
M−1·s−1; first order in NO2, first order in F2; second order overall; doubling [NO2] will double the reaction rate.

Determining the Rate Law of a Reaction


The number of fundamentally different mechanisms (sets of steps in a reaction) is actually rather small compared to the large
number of chemical reactions that can occur. Thus understanding reaction mechanisms can simplify what might seem to be a
confusing variety of chemical reactions. The first step in discovering the reaction mechanism is to determine the reaction’s rate law.
This can be done by designing experiments that measure the concentration(s) of one or more reactants or products as a function of
time. For the reaction A + B → products , for example, we need to determine k and the exponents m and n in the following
equation:
m n
rate = k[A] [B] (5.3.12)

To do this, we might keep the initial concentration of B constant while varying the initial concentration of A and calculating the
initial reaction rate. This information would permit us to deduce the reaction order with respect to A. Similarly, we could determine
the reaction order with respect to B by studying the initial reaction rate when the initial concentration of A is kept constant while
the initial concentration of B is varied. In earlier examples, we determined the reaction order with respect to a given reactant by
comparing the different rates obtained when only the concentration of the reactant in question was changed. An alternative way of
determining reaction orders is to set up a proportion using the rate laws for two different experiments. Rate data for a hypothetical
reaction of the type A + B → products are given in Table 5.3.1.
Table 5.3.1 : Rate Data for a Hypothetical Reaction of the Form A + B → products
Experiment [A] (M) [B] (M) Initial Rate (M/min)

1 0.50 0.50 8.5 × 10−3

2 0.75 0.50 19 × 10−3

3 1.00 0.50 34 × 10−3

4 0.50 0.75 8.5 × 10−3

5 0.50 1.00 8.5 × 10−3

5.3.5 https://chem.libretexts.org/@go/page/81882
The general rate law for the reaction is given in Equation 5.3.12. We can obtain m or n directly by using a proportion of the rate
laws for two experiments in which the concentration of one reactant is the same, such as Experiments 1 and 3 in Table 5.3.3.
m n
rate1 k[ A1 ] [ B1 ]
=
m n
rate3 k[ A3 ] [ B3 ]

Inserting the appropriate values from Table 5.3.3,


−3 m n
8.5 × 10  M/min k[0.50 M] [0.50 M]
=
−3 m n
34 × 10  M/min k[1.00 M] [0.50 M]

Because 1.00 to any power is 1, [1.00 M]m = 1.00 M. We can cancel like terms to give 0.25 = [0.50]m, which can also be written as
1/4 = [1/2]m. Thus we can conclude that m = 2 and that the reaction is second order in A. By selecting two experiments in which
the concentration of B is the same, we were able to solve for m.
Conversely, by selecting two experiments in which the concentration of A is the same (e.g., Experiments 5 and 1), we can solve for
n.
m n
rate1 k[ A1 ] [ B1 ]
=
m n
rate5 k[ A5 ] [ B5 ]

Substituting the appropriate values from Table 5.3.3,


−3 m n
8.5 × 10  M/min k[0.50 M] [0.50 M]
=
−3 m n
8.5 × 10  M/min k[0.50 M] [1.00 M]

Canceling leaves 1.0 = [0.50]n, which gives n =0 ; that is, the reaction is zeroth order in B . The experimentally determined rate
law is therefore
rate = k[A]2[B]0 = k[A]2
We can now calculate the rate constant by inserting the data from any row of Table 5.3.3 into the experimentally determined rate
law and solving for k . Using Experiment 2, we obtain
19 × 10−3 M/min = k(0.75 M)2
3.4 × 10−2 M−1·min−1 = k
You should verify that using data from any other row of Table 5.3.1 gives the same rate constant. This must be true as long as the
experimental conditions, such as temperature and solvent, are the same.

 Example 5.3.3
Nitric oxide is produced in the body by several different enzymes and acts as a signal that controls blood pressure, long-term
memory, and other critical functions. The major route for removing NO from biological fluids is via reaction with O to give 2

N O , which then reacts rapidly with water to give nitrous acid and nitric acid:
2

These reactions are important in maintaining steady levels of NO. The following table lists kinetics data for the reaction of NO
with O2 at 25°C:

2N O(g) + O2 (g) → 2N O2 (g)

Determine the rate law for the reaction and calculate the rate constant.
rate law for the reaction and calculate the rate constant.
Experiment [NO]0 (M) [O2]0 (M) Initial Rate (M/s)

1 0.0235 0.0125 7.98 × 10−3

2 0.0235 0.0250 15.9 × 10−3

3 0.0470 0.0125 32.0 × 10−3

5.3.6 https://chem.libretexts.org/@go/page/81882
Experiment [NO]0 (M) [O2]0 (M) Initial Rate (M/s)

4 0.0470 0.0250 63.5 × 10−3

Given: balanced chemical equation, initial concentrations, and initial rates


Asked for: rate law and rate constant

Strategy:
A. Compare the changes in initial concentrations with the corresponding changes in rates of reaction to determine the reaction
order for each species. Write the rate law for the reaction.
B. Using data from any experiment, substitute appropriate values into the rate law. Solve the rate equation for k.

Solution
A Comparing Experiments 1 and 2 shows that as [O2] is doubled at a constant value of [NO2], the reaction rate approximately
doubles. Thus the reaction rate is proportional to [O2]1, so the reaction is first order in O2. Comparing Experiments 1 and 3
shows that the reaction rate essentially quadruples when [NO] is doubled and [O2] is held constant. That is, the reaction rate is
proportional to [NO]2, which indicates that the reaction is second order in NO. Using these relationships, we can write the rate
law for the reaction:
rate = k[NO]2[O2]
B The data in any row can be used to calculate the rate constant. Using Experiment 1, for example, gives
−3
rate 7.98 × 10  M/s
3 −2 −1
k = = = 1.16 × 10 M ⋅s
2 2
[NO] [ O2 ] (0.0235 M) (0.0125 M)

Alternatively, using Experiment 2 gives


−3
rate 15.9 × 10  M/s
3 −2 −1
k = = = 1.15 × 10 M ⋅s
2 2
[NO] [ O2 ] (0.0235 M) (0.0250 M)

The difference is minor and associated with significant digits and likely experimental error in making the table.
The overall reaction order (m + n) = 3 , so this is a third-order reaction whose rate is determined by three reactants. The units
of the rate constant become more complex as the overall reaction order increases.

 Exercise 5.3.3
The peroxydisulfate ion (S2O82−) is a potent oxidizing agent that reacts rapidly with iodide ion in water:
2− − 2− −
S2 O + 3I → 2S O +I
8(aq) (aq) 4(aq) 3(aq)

The following table lists kinetics data for this reaction at 25°C. Determine the rate law and calculate the rate constant.
kinetics data for this reaction at 25°C.
Experiment [S2O82−]0 (M) [I−]0 (M) Initial Rate (M/s)

1 0.27 0.38 2.05

2 0.40 0.38 3.06

3 0.40 0.22 1.76

Answer:
rate = k[S2O82−][I−]; k = 20 M−1·s−1

5.3.7 https://chem.libretexts.org/@go/page/81882
I'm not a robot
reCAPTCHA
Privacy - Terms

About this page

Our systems have detected unusual traffic from your


computer network This page checks to see if it's really you
A Video Discussing Initial Rates and Rate Law Expressions. Video Link: Initial Rates and Rate Law Expressions(opens in new
window) [youtu.be]

Summary
The rate law for a reaction is a mathematical relationship between the reaction rate and the concentrations of species in solution.
Rate laws can be expressed either as a differential rate law, describing the change in reactant or product concentrations as a
function of time, or as an integrated rate law, describing the actual concentrations of reactants or products as a function of time. The
rate constant (k) of a rate law is a constant of proportionality between the reaction rate and the reactant concentration. The exponent
to which a concentration is raised in a rate law indicates the reaction order, the degree to which the reaction rate depends on the
concentration of a particular reactant.

5.3: Concentration and Rates (Rate Laws) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
14.3: Concentration and Rates (Differential Rate Laws) is licensed CC BY-NC-SA 3.0.

5.3.8 https://chem.libretexts.org/@go/page/81882
5.4: Determining Rate Laws from Initial Rates (Differential Rate Laws)
It is sometimes helpful to use an explicit algebraic method, often referred to as the method of initial rates, to determine the orders in
rate laws. To use this method, we select two sets of rate data that differ in the concentration of only one reactant and set up a ratio
of the two rates and the two rate laws. After canceling terms that are equal, we are left with an equation that contains only one
unknown, the coefficient of the concentration that varies. We then solve this equation for the coefficient.

etermining a Rate Law from Initial Rates


Ozone in the upper atmosphere is depleted when it reacts with nitrogen oxides. The rates of the reactions of nitrogen oxides
with ozone are important factors in deciding how significant these reactions are in the formation of the ozone hole over
Antarctica (Figure 5.4.1). One such reaction is the combination of nitric oxide, N O , with ozone, O3:

NO(g) + O (g) ⟶ NO (g) + O (g)


3 2 2

Figure 5.4.1 : Over the past several years, the atmospheric ozone concentration over Antarctica has decreased during the
winter. This map shows the decreased concentration as a purple area. The Antarctic hole is stabilizing and may be slowly
recovering. (credit: modification of work by NASA)
This reaction has been studied in the laboratory, and the following rate data were determined at 25 °C.
Δ[ NO ]
Trial [N O] (mol/L) [ O3 ] (mol/L) 2
(mol L
−1 −1
s )
Δt

1 1.00 × 10−6 3.00 × 10−6 6.60 × 10−5

2 1.00 × 10−6 6.00 × 10−6 1.32 × 10−4

3 1.00 × 10−6 9.00 × 10−6 1.98 × 10−4

4 2.00 × 10−6 9.00 × 10−6 3.96 × 10−4

5 3.00 × 10−6 9.00 × 10−6 5.94 × 10−4

Determine the rate law and the rate constant for the reaction at 25 °C.
Solution
The rate law will have the form:
m n
rate = k[NO] [O ]
3

We can determine the values of m, n, and k from the experimental data using the following three-part process:
1. Determine the value of m from the data in which [NO] varies and [O3] is constant. In the last three experiments, [NO] varies
while [O3] remains constant. When [NO] doubles from trial 3 to 4, the rate doubles, and when [NO] triples from trial 3 to 5, the
rate also triples. Thus, the rate is also directly proportional to [NO], and m in the rate law is equal to 1.
2. Determine the value of n from data in which [O3] varies and [NO] is constant. In the first three experiments, [NO] is constant
and [O3] varies. The reaction rate changes in direct proportion to the change in [O3]. When [O3] doubles from trial 1 to 2, the
rate doubles; when [O3] triples from trial 1 to 3, the rate increases also triples. Thus, the rate is directly proportional to [O3], and
n is equal to 1.The rate law is thus:

5.4.1 https://chem.libretexts.org/@go/page/81958
1 1
rate = k[NO] [ O ] = k[NO][ O ]
3 3

3. Determine the value of k from one set of concentrations and the corresponding rate.
rate
k =
[NO][ O3 ]

−5 −1 −1
6.60 × 10 mol L s
=
−6 −1 −6 −1
(1.00 × 10 mol L )(3.00 × 10 mol L )

7 −1 −1
= 2.20 × 10 L mo l s

The large value of k tells us that this is a fast reaction that could play an important role in ozone depletion if [NO] is large
enough.

Exercise 5.4.1

Acetaldehyde decomposes when heated to yield methane and carbon monoxide according to the equation:
CH CHO(g) ⟶ CH (g) + CO(g)
3 4

Determine the rate law and the rate constant for the reaction from the following experimental data:
Δ[ CH CHO]
Trial [CH3CHO] (mol/L) −
3
(mol L
−1 −1
s )
Δt

1 1.75 × 10−3 2.06 × 10−11

2 3.50 × 10−3 8.24 × 10−11

3 7.00 × 10−3 3.30 × 10−10

Answer
rate = k[ CH CHO]
3
2
with k = 6.73 × 10−6 L/mol/s

Determining Rate Laws from Initial Rates


Using the initial rates method and the experimental data, determine the rate law and the value of the rate constant for this
reaction:

2 NO(g) + Cl (g) ⟶ 2 NOCl(g)


2

Δ[NO]
Trial [NO] (mol/L) [Cl2] (mol/L) − (mol L
−1 −1
s )
Δt

1 0.10 0.10 0.00300

2 0.10 0.15 0.00450

3 0.15 0.10 0.00675

Solution
The rate law for this reaction will have the form:
m n
rate = k[NO] [ Cl ]
2

As in Example 5.4.21, we can approach this problem in a stepwise fashion, determining the values of m and n from the
experimental data and then using these values to determine the value of k. In this example, however, we will use a different
approach to determine the values of m and n:
1. Determine the value of m from the data in which [NO] varies and [Cl2] is constant. We can write the ratios with the subscripts x
and y to indicate data from two different trials:

5.4.2 https://chem.libretexts.org/@go/page/81958
m n
ratex k[NO]x [ Cl ]x
2
=
m n
ratey k[NO]y [ Cl ]y
2

Using the third trial and the first trial, in which [Cl2] does not vary, gives:
m n
rate 3 0.00675 k(0.15 ) (0.10 )
= =
m n
rate 1 0.00300 k(0.10 ) (0.10 )

After canceling equivalent terms in the numerator and denominator, we are left with:
m
0.00675 (0.15)
=
m
0.00300 (0.10)

which simplifies to:


m
2.25 = (1.5)

We can use natural logs to determine the value of the exponent m:


ln(2.25)
ln(2.25) = m ln(1.5) = m2 = m (5.4.1)
ln(1.5)

We can confirm the result easily, since:


2
1.5 = 2.25

2. Determine the value of n from data in which [Cl2] varies and [NO] is constant.
m n
rate 2 0.00450 k(0.10 ) (0.15 )
= =
m n
rate 1 0.00300 k(0.10 ) (0.10 )

Cancelation gives:
n
0.0045 (0.15)
=
n
0.0030 (0.10)

which simplifies to:


n
1.5 = (1.5)

Thus n must be 1, and the form of the rate law is:


m n 2
Rate = k[NO] [ Cl ] = k[NO] [ Cl ]
2 2

3. Determine the numerical value of the rate constant k with appropriate units. The units for the rate of a reaction are mol/L/s. The
units for k are whatever is needed so that substituting into the rate law expression affords the appropriate units for the rate. In
this example, the concentration units are mol3/L3. The units for k should be mol−2 L2/s so that the rate is in terms of mol/L/s.
To determine the value of k once the rate law expression has been solved, simply plug in values from the first experimental trial
and solve for k:
−1 −1 −1 2 −1 1
0.00300 mol L s = k(0.10 mol L ) (0.10 mol L )

−2 2 −1
k = 3.0 mol L s

Exercise 5.4.2

Use the provided initial rate data to derive the rate law for the reaction whose equation is:
− − − −
OCl (aq) + I (aq) ⟶ OI (aq) + Cl (aq) (5.4.2)

Trial [OCl−] (mol/L) [I−] (mol/L) Initial Rate (mol/L/s)

1 0.0040 0.0020 0.00184

2 0.0020 0.0040 0.00092

5.4.3 https://chem.libretexts.org/@go/page/81958
Trial [OCl−] (mol/L) [I−] (mol/L) Initial Rate (mol/L/s)

3 0.0020 0.0020 0.00046

Determine the rate law expression and the value of the rate constant k with appropriate units for this reaction.

Answer
x y
rate 2 0.00092 k(0.0020 ) (0.0040 )
= =
x y
rate 3 0.00046 k(0.0020 ) (0.0020 )

y
2.00 = 2.00

so

y =1

Now the other reactant


x y
rate 1 0.00184 k(0.0040 ) (0.0020 )
= =
x y
rate 2 0.00092 k(0.0020 ) (0.0040 )

x
2
2.00 =
y
2
x
2
2.00 =
1
2
x
4.00 = 2

x =2

Substituting the concentration data from trial 1 and solving for k yields:
− 2 − 1
rate = k[ OCl ] [I ]

2 1
0.00184 = k(0.0040 ) (0.0020 )

4 −2 2 −1
k = 5.75 × 10 mo l L s

Contributors and Attributions


Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).

5.4: Determining Rate Laws from Initial Rates (Differential Rate Laws) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed,
and/or curated by LibreTexts.

5.4.4 https://chem.libretexts.org/@go/page/81958
5.5: The Change of Concentration with Time (Integrated Rate Laws)
 Learning Objectives
To apply rate laws to zeroth, first and second order reactions.

Either the differential rate law or the integrated rate law can be used to determine the reaction order from experimental data. Often,
the exponents in the rate law are the positive integers: 1 and 2 or even 0. Thus the reactions are zeroth, first, or second order in each
reactant. The common patterns used to identify the reaction order are described in this section, where we focus on characteristic
types of differential and integrated rate laws and how to determine the reaction order from experimental data. The learning
objective of this Module is to know how to determine the reaction order from experimental data.

Zeroth-Order Reactions
A zeroth-order reaction is one whose rate is independent of concentration; its differential rate law is

rate = k.

We refer to these reactions as zeroth order because we could also write their rate in a form such that the exponent of the reactant in
the rate law is 0:
Δ[A]
0
rate = − = k[reactant] = k(1) = k (5.5.1)
Δt

Because rate is independent of reactant concentration, a graph of the concentration of any reactant as a function of time is a straight
line with a slope of −k . The value of k is negative because the concentration of the reactant decreases with time. Conversely, a
graph of the concentration of any product as a function of time is a straight line with a slope of k , a positive value.

Figure The graph of a zeroth-order reaction. The change in concentration of reactant and
5.5.1 :
product with time produces a straight line.
Graph of concentration against time. The reactant is in purple and has a slope of minus k. The product is in green and has a slope of
positive k.
The integrated rate law for a zeroth-order reaction also produces a straight line and has the general form
[A] = [A]0 − kt (5.5.2)

where [A] is the initial concentration of reactant A . Equation 5.5.2 has the form of the algebraic equation for a straight line,
0

y = mx + b,

with y = [A] , mx = −kt , and b = [A] .) 0

 Units
In a zeroth-order reaction, the rate constant must have the same units as the reaction rate, typically moles per liter per second.

Although it may seem counterintuitive for the reaction rate to be independent of the reactant concentration(s), such reactions are
rather common. They occur most often when the reaction rate is determined by available surface area. An example is the

5.5.1 https://chem.libretexts.org/@go/page/81883
decomposition of N2O on a platinum (Pt) surface to produce N2 and O2, which occurs at temperatures ranging from 200°C to
400°C:
Pt

2 N2 O(g) −→ 2 N2 (g) + O2 (g) (5.5.3)

Without a platinum surface, the reaction requires temperatures greater than 700°C, but between 200°C and 400°C, the only factor
that determines how rapidly N2O decomposes is the amount of Pt surface available (not the amount of Pt). As long as there is
enough N2O to react with the entire Pt surface, doubling or quadrupling the N2O concentration will have no effect on the reaction
rate. At very low concentrations of N2O, where there are not enough molecules present to occupy the entire available Pt surface,
the reaction rate is dependent on the N2O concentration. The reaction rate is as follows:
1 Δ[ N2 O] 1 Δ[ N2 ] Δ[ O2 ]
0
rate = − ( ) = ( ) = = k[ N2 O] =k (5.5.4)
2 Δt 2 Δt Δt

Thus the rate at which N2O is consumed and the rates at which N2 and O2 are produced are independent of concentration. As
shown in Figure 5.5.2, the change in the concentrations of all species with time is linear. Most important, the exponent (0)
corresponding to the N2O concentration in the experimentally derived rate law is not the same as the reactant’s stoichiometric
coefficient in the balanced chemical equation (2). For this reaction, as for all others, the rate law must be determined
experimentally.

Figure 5.5.2 : A Zeroth-Order Reaction. This graph shows the concentrations of reactants and products versus time for the zeroth-
order catalyzed decomposition of N2O to N2 and O2 on a Pt surface. The change in the concentrations of all species with time is
linear.
Graph of concentration against time. N2O is the reactiant is graphed in purple. O2 is one of the products and is graphed in green.
The second product is N2 which is graphed in red
A zeroth-order reaction that takes place in the human liver is the oxidation of ethanol (from alcoholic beverages) to acetaldehyde,
catalyzed by the enzyme alcohol dehydrogenase. At high ethanol concentrations, this reaction is also a zeroth-order reaction. The
overall reaction equation is
CH3CH2OH and <span class=
NAD
plus react with alcohol dehydrogenase to produce CH3COH,
NADH
, and H plus." style="width: 816px; height: 76px;" width="816px" height="76px" data-cke-saved-
src="/@api/deki/files/15941/14.9.jpg" src="/@api/deki/files/15941/14.9.jpg" data-quail-id="61">
where NAD+ (nicotinamide adenine dinucleotide) and NADH (reduced nicotinamide adenine dinucleotide) are the oxidized and
reduced forms, respectively, of a species used by all organisms to transport electrons. When an alcoholic beverage is consumed, the
ethanol is rapidly absorbed into the blood. Its concentration then decreases at a constant rate until it reaches zero (Figure 5.5.3a).
An average 70 kg person typically takes about 2.5 h to oxidize the 15 mL of ethanol contained in a single 12 oz can of beer, a 5 oz
glass of wine, or a shot of distilled spirits (such as whiskey or brandy). The actual rate, however, varies a great deal from person to
person, depending on body size and the amount of alcohol dehydrogenase in the liver. The reaction rate does not increase if a
greater quantity of alcohol is consumed over the same period of time because the reaction rate is determined only by the amount of
enzyme present in the liver. Contrary to popular belief, the caffeine in coffee is ineffective at catalyzing the oxidation of ethanol.

5.5.2 https://chem.libretexts.org/@go/page/81883
When the ethanol has been completely oxidized and its concentration drops to essentially zero, the rate of oxidation also drops
rapidly (part (b) in Figure 5.5.3).

Figure 5.5.3 : The Catalyzed Oxidation of Ethanol (a) The concentration of ethanol in human blood decreases linearly with time,
which is typical of a zeroth-order reaction. (b) The rate at which ethanol is oxidized is constant until the ethanol concentration
reaches essentially zero, at which point the reaction rate drops to zero.
These examples illustrate two important points:
1. In a zeroth-order reaction, the reaction rate does not depend on the reactant concentration.
2. A linear change in concentration with time is a clear indication of a zeroth-order reaction.

I'm not a robot


reCAPTCHA
Privacy - Terms

About this page

Our systems have detected unusual traffic from your


computer network This page checks to see if it's really you
A Video Discussing Zero-Order Reactions. Video Link: Zero-Order Reactions(opens in new window) [youtu.be] (opens in new
window)

First-Order Reactions
In a first-order reaction, the reaction rate is directly proportional to the concentration of one of the reactants. First-order reactions
often have the general form A → products. The differential rate for a first-order reaction is as follows:
Δ[A]
rate = − = k[A] (5.5.5)
Δt

If the concentration of A is doubled, the reaction rate doubles; if the concentration of A is increased by a factor of 10, the reaction
rate increases by a factor of 10, and so forth. Because the units of the reaction rate are always moles per liter per second, the units
of a first-order rate constant are reciprocal seconds (s−1).
The integrated rate law for a first-order reaction can be written in two different ways: one using exponents and one using
logarithms. The exponential form is as follows:
−kt
[A] = [A]0 e (5.5.6)

5.5.3 https://chem.libretexts.org/@go/page/81883
where [A] is the initial concentration of reactant A at t = 0 ; k is the rate constant; and e is the base of the natural logarithms,
0

which has the value 2.718 to three decimal places. Recall that an integrated rate law gives the relationship between reactant
concentration and time. Equation 5.5.6 predicts that the concentration of A will decrease in a smooth exponential curve over time.
By taking the natural logarithm of each side of Equation 5.5.6 and rearranging, we obtain an alternative logarithmic expression of
the relationship between the concentration of A and t :
ln[A] = ln[A]0 − kt (5.5.7)

Because Equation 5.5.7 has the form of the algebraic equation for a straight line,

y = mx + b,

with y = ln[A] and b = ln[A] , a plot of ln[A] versus t for a first-order reaction should give a straight line with a slope of −k and
0

an intercept of ln[A] . Either the differential rate law (Equation 5.5.5) or the integrated rate law (Equation 5.5.7) can be used to
0

determine whether a particular reaction is first order.

Figure 5.5.4 : Graphs of a first-order reaction. The expected shapes of the curves for plots of reactant concentration versus time
(top) and the natural logarithm of reactant concentration versus time (bottom) for a first-order reaction.

I'm not a robot


reCAPTCHA
Privacy - Terms

About this page

Our systems have detected unusual traffic from your


computer network This page checks to see if it's really you
Video Discussing the The First-Order Integrated Rate Law Equation: The First-Order Integrated Rate Law Equation(opens in new
window) [youtu.be]

First-order reactions are very common. One reaction that exhibits apparent first-order kinetics is the hydrolysis of the anticancer
drug cisplatin. Cisplatin, the first “inorganic” anticancer drug to be discovered, is unique in its ability to cause complete remission
of the relatively rare, but deadly cancers of the reproductive organs in young adults. The structures of cisplatin and its hydrolysis
product are as follows:

Figure 5.5.5 : Cis-platin reaction with water.

5.5.4 https://chem.libretexts.org/@go/page/81883
Both platinum compounds have four groups arranged in a square plane around a Pt(II) ion. The reaction shown in Figure 5.5.5 is
important because cisplatin, the form in which the drug is administered, is not the form in which the drug is active. Instead, at least
one chloride ion must be replaced by water to produce a species that reacts with deoxyribonucleic acid (DNA) to prevent cell
division and tumor growth. Consequently, the kinetics of the reaction in Figure 5.5.4 have been studied extensively to find ways of
maximizing the concentration of the active species.

If a plot of reactant concentration versus time is not linear but a plot of the natural
logarithm of reactant concentration versus time is linear, then the reaction is first order.
The rate law and reaction order of the hydrolysis of cisplatin are determined from experimental data, such as those displayed in
Table 5.5.1. The table lists initial rate data for four experiments in which the reaction was run at pH 7.0 and 25°C but with different
initial concentrations of cisplatin.
Table 5.5.1 : Rates of Hydrolysis of Cisplatin as a Function of Concentration at pH 7.0 and 25°C
Experiment [Cisplatin]0 (M) Initial Rate (M/min)

1 0.0060 9.0 × 10−6

2 0.012 1.8 × 10−5

3 0.024 3.6 × 10−5

4 0.030 4.5 × 10−5

Because the reaction rate increases with increasing cisplatin concentration, we know this cannot be a zeroth-order reaction.
Comparing Experiments 1 and 2 in Table 5.5.1 shows that the reaction rate doubles [(1.8 × 10−5 M/min) ÷ (9.0 × 10−6 M/min) =
2.0] when the concentration of cisplatin is doubled (from 0.0060 M to 0.012 M). Similarly, comparing Experiments 1 and 4 shows
that the reaction rate increases by a factor of 5 [(4.5 × 10−5 M/min) ÷ (9.0 × 10−6 M/min) = 5.0] when the concentration of cisplatin
is increased by a factor of 5 (from 0.0060 M to 0.030 M). Because the reaction rate is directly proportional to the concentration of
the reactant, the exponent of the cisplatin concentration in the rate law must be 1, so the rate law is rate = k[cisplatin]1. Thus the
reaction is first order. Knowing this, we can calculate the rate constant using the differential rate law for a first-order reaction and
the data in any row of Table 5.5.1. For example, substituting the values for Experiment 3 into Equation 5.5.5,
3.6 × 10−5 M/min = k(0.024 M)
1.5 × 10−3 min−1 = k
Knowing the rate constant for the hydrolysis of cisplatin and the rate constants for subsequent reactions that produce species that
are highly toxic enables hospital pharmacists to provide patients with solutions that contain only the desired form of the drug.

 Example 5.5.1

At high temperatures, ethyl chloride produces HCl and ethylene by the following reaction:
Δ

CH CH Cl(g) −
→ HCl(g) + C H (g)
3 2 2 4

Using the rate data for the reaction at 650°C presented in the following table, calculate the reaction order with respect to the
concentration of ethyl chloride and determine the rate constant for the reaction.
data for the reaction at 650°C
Experiment [CH3CH2Cl]0 (M) Initial Rate (M/s)

1 0.010 1.6 × 10−8

2 0.015 2.4 × 10−8

3 0.030 4.8 × 10−8

4 0.040 6.4 × 10−8

Given: balanced chemical equation, initial concentrations of reactant, and initial rates of reaction
Asked for: reaction order and rate constant

5.5.5 https://chem.libretexts.org/@go/page/81883
Strategy:
A. Compare the data from two experiments to determine the effect on the reaction rate of changing the concentration of a
species.
B. Compare the observed effect with behaviors characteristic of zeroth- and first-order reactions to determine the reaction
order. Write the rate law for the reaction.
C Use measured concentrations and rate data from any of the experiments to find the rate constant.

Solution
The reaction order with respect to ethyl chloride is determined by examining the effect of changes in the ethyl chloride
concentration on the reaction rate.
A Comparing Experiments 2 and 3 shows that doubling the concentration doubles the reaction rate, so the reaction rate is
proportional to [CH3CH2Cl]. Similarly, comparing Experiments 1 and 4 shows that quadrupling the concentration quadruples
the reaction rate, again indicating that the reaction rate is directly proportional to [CH3CH2Cl].
B This behavior is characteristic of a first-order reaction, for which the rate law is rate = k[CH3CH2Cl].
C We can calculate the rate constant (k) using any row in the table. Selecting Experiment 1 gives the following:
1.60 × 10−8 M/s = k(0.010 M)
1.6 × 10−6 s−1 = k

 Exercise 5.5.1

Sulfuryl chloride (SO2Cl2) decomposes to SO2 and Cl2 by the following reaction:

S O2 C l2 (g) → S O2 (g) + C l2 (g)

Data for the reaction at 320°C are listed in the following table. Calculate the reaction order with regard to sulfuryl chloride and
determine the rate constant for the reaction.
Data for the reaction at 320°C
Experiment [SO2Cl2]0 (M) Initial Rate (M/s)

1 0.0050 1.10 × 10−7

2 0.0075 1.65 × 10−7

3 0.0100 2.20 × 10−7

4 0.0125 2.75 × 10−7

Answer
first order; k = 2.2 × 10−5 s−1

We can also use the integrated rate law to determine the reaction rate for the hydrolysis of cisplatin. To do this, we examine the
change in the concentration of the reactant or the product as a function of time at a single initial cisplatin concentration. Figure
5.5.6a shows plots for a solution that originally contained 0.0100 M cisplatin and was maintained at pH 7 and 25°C.

5.5.6 https://chem.libretexts.org/@go/page/81883
Figure 5.5.6 : The Hydrolysis of Cisplatin, a First-Order Reaction. These plots show hydrolysis of cisplatin at pH 7.0 and 25°C as
(a) the experimentally determined concentrations of cisplatin and chloride ions versus time and (b) the natural logarithm of the
cisplatin concentration versus time. The straight line in (b) is expected for a first-order reaction.
The concentration of cisplatin decreases smoothly with time, and the concentration of chloride ion increases in a similar way. When
we plot the natural logarithm of the concentration of cisplatin versus time, we obtain the plot shown in part (b) in Figure 5.5.6. The
straight line is consistent with the behavior of a system that obeys a first-order rate law. We can use any two points on the line to
calculate the slope of the line, which gives us the rate constant for the reaction. Thus taking the points from part (a) in Figure 5.5.6
for t = 100 min ([cisplatin] = 0.0086 M) and t = 1000 min ([cisplatin] = 0.0022 M),
ln[cisplatin]1000 − ln[cisplatin]100
slope =
1000 min − 100 min

ln 0.0022 − ln 0.0086 −6.12 − (−4.76)


−3 −1
−k = = = −1.51 × 10 mi n
1000 min − 100 min 900 min

−3 −1
k = 1.5 × 10 mi n

The slope is negative because we are calculating the rate of disappearance of cisplatin. Also, the rate constant has units of min−1
because the times plotted on the horizontal axes in parts (a) and (b) in Figure 5.5.6 are in minutes rather than seconds.
The reaction order and the magnitude of the rate constant we obtain using the integrated rate law are exactly the same as those we
calculated earlier using the differential rate law. This must be true if the experiments were carried out under the same conditions.

I'm not a robot


reCAPTCHA
Privacy - Terms

About this page

Our systems have detected unusual traffic from your


computer network This page checks to see if it's really you
Video Example Using the First-Order Integrated Rate Law Equation:

Example Using the First-Order Integrated Rate Law Equation(opens in new window) [youtu.be]

5.5.7 https://chem.libretexts.org/@go/page/81883
 Example 5.5.2

If a sample of ethyl chloride with an initial concentration of 0.0200 M is heated at 650°C, what is the concentration of ethyl
chloride after 10 h? How many hours at 650°C must elapse for the concentration to decrease to 0.0050 M (k = 1.6 × 10−6 s−1) ?
Given: initial concentration, rate constant, and time interval
Asked for: concentration at specified time and time required to obtain particular concentration

Strategy:
A. Substitute values for the initial concentration ([A]0) and the calculated rate constant for the reaction (k) into the integrated
rate law for a first-order reaction. Calculate the concentration ([A]) at the given time t.
B. Given a concentration [A], solve the integrated rate law for time t.

Solution
The exponential form of the integrated rate law for a first-order reaction (Equation 5.5.6) is [A] = [A]0e−kt.
A Having been given the initial concentration of ethyl chloride ([A]0) and having the rate constant of k = 1.6 × 10−6 s−1, we can
use the rate law to calculate the concentration of the reactant at a given time t. Substituting the known values into the integrated
rate law,
−kt
[C H3 C H2 Cl ]10 h = [C H3 C H2 Cl ]0 e

−6 −1
−(1.6× 10  s )[(10 h)(60 min/h)(60 s/min)]
= 0.0200 M(e )

= 0.0189 M

We could also have used the logarithmic form of the integrated rate law (Equation 5.5.7):

ln[C H3 C H2 Cl ]10 h = ln[C H3 C H2 Cl ]0 − kt

−6 −1
= ln 0.0200 − (1.6 × 10  s )[(10 h)(60 min/h)(60 s/min)]

= −3.912 − 0.0576 = −3.970

−3.970
[C H3 C H2 Cl ]10 h =e  M

= 0.0189 M

B To calculate the amount of time required to reach a given concentration, we must solve the integrated rate law for t .
Equation 5.5.7 gives the following:
ln[C H3 C H2 Cl ]t = ln[C H3 C H2 Cl ]0 − kt

[C H3 C H2 Cl ]0
kt = ln[C H3 C H2 Cl ]0 − ln[C H3 C H2 Cl ]t = ln
[C H3 C H2 Cl ]t

1 [C H3 C H2 Cl ]0 1 0.0200 M
t = (ln ) = (ln )
−6 −1
k [C H3 C H2 Cl ]t 1.6 × 10  s 0.0050 M

ln 4.0
5 2
= = 8.7 × 10  s = 240 h = 2.4 × 10  h
−6 −1
1.6 × 10  s

5.5.8 https://chem.libretexts.org/@go/page/81883
 Exercise 5.5.2
In the exercise in Example 5.5.1, you found that the decomposition of sulfuryl chloride (SO 2
Cl
2
) is first order, and you
calculated the rate constant at 320°C.
a. Use the form(s) of the integrated rate law to find the amount of SO 2
Cl
2
that remains after 20 h if a sample with an original
concentration of 0.123 M is heated at 320°C.
b. How long would it take for 90% of the SO2Cl2 to decompose?

Answer a
0.0252 M
Answer b
29 h

Second-Order Reactions
The simplest kind of second-order reaction is one whose rate is proportional to the square of the concentration of one reactant.
These generally have the form

2 A → products⋅

A second kind of second-order reaction has a reaction rate that is proportional to the product of the concentrations of two reactants.
Such reactions generally have the form A + B → products. An example of the former is a dimerization reaction, in which two
smaller molecules, each called a monomer, combine to form a larger molecule (a dimer).
The differential rate law for the simplest second-order reaction in which 2A → products is as follows:
Δ[A]
2
rate = − = k[A] (5.5.8)
2Δt

Consequently, doubling the concentration of A quadruples the reaction rate. For the units of the reaction rate to be moles per liter
per second (M/s), the units of a second-order rate constant must be the inverse (M−1·s−1). Because the units of molarity are
expressed as mol/L, the unit of the rate constant can also be written as L(mol·s).
For the reaction 2A → products, the following integrated rate law describes the concentration of the reactant at a given time:
1 1
= + kt (5.5.9)
[A] [A]0

Because Equation 5.5.9 has the form of an algebraic equation for a straight line, y = mx + b, with y = 1/[A] and b = 1/[A]0, a plot of
1/[A] versus t for a simple second-order reaction is a straight line with a slope of k and an intercept of 1/[A]0.

Second-order reactions generally have the form 2A → products or A + B → products.

I'm not a robot


reCAPTCHA
Privacy - Terms

About this page

Our systems have detected unusual traffic from your


computer network This page checks to see if it's really you

5.5.9 https://chem.libretexts.org/@go/page/81883
Video Discussing the Second-Order Integrated Rate Law Equation: Second-Order Integrated Rate Law Equation(opens in new
window) [youtu.be]

Simple second-order reactions are common. In addition to dimerization reactions, two other examples are the decomposition of
NO2 to NO and O2 and the decomposition of HI to I2 and H2. Most examples involve simple inorganic molecules, but there are
organic examples as well. We can follow the progress of the reaction described in the following paragraph by monitoring the
decrease in the intensity of the red color of the reaction mixture.
Many cyclic organic compounds that contain two carbon–carbon double bonds undergo a dimerization reaction to give complex
structures. One example is as follows:

Figure 5.5.7
For simplicity, we will refer to this reactant and product as “monomer” and “dimer,” respectively. The systematic name of the
monomer is 2,5-dimethyl-3,4-diphenylcyclopentadienone. The systematic name of the dimer is the name of the monomer followed
by “dimer.” Because the monomers are the same, the general equation for this reaction is 2A → product. This reaction represents
an important class of organic reactions used in the pharmaceutical industry to prepare complex carbon skeletons for the synthesis of
drugs. Like the first-order reactions studied previously, it can be analyzed using either the differential rate law (Equation 5.5.8) or
the integrated rate law (Equation 5.5.9).
Table 5.5.2 : Rates of Reaction as a Function of Monomer Concentration for an Initial Monomer Concentration of 0.0054 M
Time (min) [Monomer] (M) Instantaneous Rate (M/min)

10 0.0044 8.0 × 10−5

26 0.0034 5.0 × 10−5

44 0.0027 3.1 × 10−5

70 0.0020 1.8 × 10−5

120 0.0014 8.0 × 10−6

To determine the differential rate law for the reaction, we need data on how the reaction rate varies as a function of monomer
concentrations, which are provided in Table 5.5.2. From the data, we see that the reaction rate is not independent of the monomer
concentration, so this is not a zeroth-order reaction. We also see that the reaction rate is not proportional to the monomer
concentration, so the reaction is not first order. Comparing the data in the second and fourth rows shows that the reaction rate
decreases by a factor of 2.8 when the monomer concentration decreases by a factor of 1.7:
−5 −3
5.0 × 10  M/min 3.4 × 10  M
= 2.8 and = 1.7
−5 −3
1.8 × 10  M/min 2.0 × 10  M

Because (1.7)2 = 2.9 ≈ 2.8, the reaction rate is approximately proportional to the square of the monomer concentration.
rate ∝ [monomer]2
This means that the reaction is second order in the monomer. Using Equation 5.5.8 and the data from any row in Table 5.5.2, we
can calculate the rate constant. Substituting values at time 10 min, for example, gives the following:

5.5.10 https://chem.libretexts.org/@go/page/81883
2
rate = k[A] (5.5.10)

−5 −3 2
8.0 × 10  M/min = k(4.4 × 10  M) (5.5.11)
−1 −1
4.1  M ⋅ min =k (5.5.12)

We can also determine the reaction order using the integrated rate law. To do so, we use the decrease in the concentration of the
monomer as a function of time for a single reaction, plotted in Figure 5.5.8a. The measurements show that the concentration of the
monomer (initially 5.4 × 10−3 M) decreases with increasing time. This graph also shows that the reaction rate decreases smoothly
with increasing time. According to the integrated rate law for a second-order reaction, a plot of 1/[monomer] versus t should be a
straight line, as shown in Figure 5.5.8b. Any pair of points on the line can be used to calculate the slope, which is the second-order
rate constant. In this example, k = 4.1 M−1·min−1, which is consistent with the result obtained using the differential rate equation.
Although in this example the stoichiometric coefficient is the same as the reaction order, this is not always the case. The reaction
order must always be determined experimentally.

Figure 5.5.8 : Dimerization of a Monomeric Compound, a Second-Order Reaction. These plots correspond to dimerization of the
monomer in Figure 14.4.6 as (a) the experimentally determined concentration of monomer versus time and (b) 1/[monomer] versus
time. The straight line in (b) is expected for a simple second-order reaction.
For two or more reactions of the same order, the reaction with the largest rate constant is the fastest. Because the units of the rate
constants for zeroth-, first-, and second-order reactions are different, however, we cannot compare the magnitudes of rate constants
for reactions that have different orders.

 Example 5.5.3
At high temperatures, nitrogen dioxide decomposes to nitric oxide and oxygen.
Δ

2NO2 (g) −
→ 2NO(g) + O2 (g)

Experimental data for the reaction at 300°C and four initial concentrations of NO2 are listed in the following table:
Experimental data for the reaction at 300°C and four initial concentrations of NO2
Experiment [NO2]0 (M) Initial Rate (M/s)

1 0.015 1.22 × 10−4

2 0.010 5.40 × 10−5

3 0.0080 3.46 × 10−5

4 0.0050 1.35 × 10−5

Determine the reaction order and the rate constant.


Given: balanced chemical equation, initial concentrations, and initial rates
Asked for: reaction order and rate constant

Strategy:

5.5.11 https://chem.libretexts.org/@go/page/81883
A. From the experiments, compare the changes in the initial reaction rates with the corresponding changes in the initial
concentrations. Determine whether the changes are characteristic of zeroth-, first-, or second-order reactions.
B. Determine the appropriate rate law. Using this rate law and data from any experiment, solve for the rate constant (k).

Solution
A We can determine the reaction order with respect to nitrogen dioxide by comparing the changes in NO2 concentrations with
the corresponding reaction rates. Comparing Experiments 2 and 4, for example, shows that doubling the concentration
quadruples the reaction rate [(5.40 × 10−5) ÷ (1.35 × 10−5) = 4.0], which means that the reaction rate is proportional to [NO2]2.
Similarly, comparing Experiments 1 and 4 shows that tripling the concentration increases the reaction rate by a factor of 9,
again indicating that the reaction rate is proportional to [NO2]2. This behavior is characteristic of a second-order reaction.
B We have rate = k[NO2]2. We can calculate the rate constant (k) using data from any experiment in the table. Selecting
Experiment 2, for example, gives the following:
2
rate = k[NO2 ]
−5 2
5.40 × 10  M/s = k(0.010 M)

−1 −1
0.54 M ⋅s =k

 Exercise 5.5.3

When the highly reactive species HO2 forms in the atmosphere, one important reaction that then removes it from the
atmosphere is as follows:

2H O2(g) → H2 O2(g) + O2(g)

The kinetics of this reaction have been studied in the laboratory, and some initial rate data at 25°C are listed in the following
table:
Some initial rate data at 25°C
Experiment [HO2]0 (M) Initial Rate (M/s)

1 1.1 × 10−8 1.7 × 10−7

2 2.5 × 10−8 8.8 × 10−7

3 3.4 × 10−8 1.6 × 10−6

4 5.0 × 10−8 3.5 × 10−6

Determine the reaction order and the rate constant.

Answer
second order in HO2; k = 1.4 × 109 M−1·s−1

If a plot of reactant concentration versus time is not linear, but a plot of 1/reactant concentration versus time is linear, then the
reaction is second order.

 Example 5.5.4
If a flask that initially contains 0.056 M NO2 is heated at 300°C, what will be the concentration of NO2 after 1.0 h? How long
will it take for the concentration of NO2 to decrease to 10% of the initial concentration? Use the integrated rate law for a
second-order reaction (Equation 5.5.9) and the rate constant calculated above.
Given: balanced chemical equation, rate constant, time interval, and initial concentration
Asked for: final concentration and time required to reach specified concentration

5.5.12 https://chem.libretexts.org/@go/page/81883
Strategy:
A. Given k, t, and [A]0, use the integrated rate law for a second-order reaction to calculate [A].
B. Setting [A] equal to 1/10 of [A]0, use the same equation to solve for t .

Solution
A We know k and [NO2]0, and we are asked to determine [NO2] at t = 1 h (3600 s). Substituting the appropriate values into
Equation 5.5.9,
1 1
= + kt
[NO2 ]3600 [NO2 ]0

1 −1 −1
= + [(0.54 M ⋅s )(3600 s)]
0.056 M

3 −1
= 2.0 × 10  M

Thus [NO2]3600 = 5.1 × 10−4 M.


B In this case, we know k and [NO2]0, and we are asked to calculate at what time [NO2] = 0.1[NO2]0 = 0.1(0.056 M) = 0.0056
M. To do this, we solve Equation 5.5.9 for t, using the concentrations given.
(1/[NO2 ]) − (1/[NO2 ]0 )
t =
k

(1/0.0056 M) − (1/0.056 M)
=
0.54 M−1 ⋅ s−1

2
= 3.0 × 10  s = 5.0 min

NO2 decomposes very rapidly; under these conditions, the reaction is 90% complete in only 5.0 min.

 Exercise 5.5.4

In the previous exercise, you calculated the rate constant for the decomposition of HO2 as k = 1.4 × 109 M−1·s−1. This high rate
constant means that HO2 decomposes rapidly under the reaction conditions given in the problem. In fact, the HO2 molecule is
so reactive that it is virtually impossible to obtain in high concentrations. Given a 0.0010 M sample of HO2, calculate the
concentration of HO2 that remains after 1.0 h at 25°C. How long will it take for 90% of the HO2 to decompose? Use the
integrated rate law for a second-order reaction (Equation 5.5.9) and the rate constant calculated in the exercise in Example
5.5.3.

Answer
2.0 × 10−13 M; 6.4 × 10−6 s

In addition to the simple second-order reaction and rate law we have just described, another very common second-order reaction
has the general form A + B → products , in which the reaction is first order in A and first order in B . The differential rate law for
this reaction is as follows:
Δ[A] Δ[B]
rate = − =− = k[A][B] (5.5.13)
Δt Δt

Because the reaction is first order both in A and in B, it has an overall reaction order of 2. (The integrated rate law for this reaction
is rather complex, so we will not describe it.) We can recognize second-order reactions of this sort because the reaction rate is
proportional to the concentrations of each reactant.

Summary
The reaction rate of a zeroth-order reaction is independent of the concentration of the reactants. The reaction rate of a first-order
reaction is directly proportional to the concentration of one reactant. The reaction rate of a simple second-order reaction is

5.5.13 https://chem.libretexts.org/@go/page/81883
proportional to the square of the concentration of one reactant. Knowing the rate law of a reaction gives clues to the reaction
mechanism.
zeroth-order reaction:
Δ[A]
rate = − =k
Δt

[A] = [A]0 − kt

first-order reaction:
Δ[A]
rate = − = k[A]
Δt

−kt
[A] = [A]0 e

ln[A] = ln[A]0 − kt

second-order reaction:
Δ[A]
2
rate = − = k[A]
Δt

1 1
= + kt
[A] [A]0

5.5: The Change of Concentration with Time (Integrated Rate Laws) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed,
and/or curated by LibreTexts.
14.4: The Change of Concentration with Time (Integrated Rate Laws) is licensed CC BY-NC-SA 3.0.

5.5.14 https://chem.libretexts.org/@go/page/81883
5.6: Using Graphs to Determine (Integrated) Rate Laws
 Learning Objectives
To use graphs to analyze the kinetics of a reaction.

You learned that the integrated rate law for each common type of reaction (zeroth, first, or second order in a single reactant) can be
plotted as a straight line. Using these plots offers an alternative to the methods described for showing how reactant concentration
changes with time and determining reaction order. We will illustrate the use of these graphs by considering the thermal
decomposition of NO2 gas at elevated temperatures, which occurs according to the following reaction:
Δ

2NO2 (g) −
→ 2NO(g) + O2 (g) (5.6.1)

Experimental data for this reaction at 330°C are listed in Table 5.6.1; they are provided as [NO2], ln[NO2], and 1/[NO2] versus
time to correspond to the integrated rate laws for zeroth-, first-, and second-order reactions, respectively.
Table 5.6.1 : Concentration of NO2 as a Function of Time at 330°C
Time (s) [NO2] (M) ln[NO2] 1/[NO2] (M−1)

0 1.00 × 10−2 −4.605 100

60 6.83 × 10−3 −4.986 146

120 5.18 × 10−3 −5.263 193

180 4.18 × 10−3 −5.477 239

240 3.50 × 10−3 −5.655 286

300 3.01 × 10−3 −5.806 332

360 2.64 × 10−3 −5.937 379

The actual concentrations of NO2 are plotted versus time in part (a) in Figure 5.6.1. Because the plot of [NO2] versus t is not a
straight line, we know the reaction is not zeroth order in NO2. A plot of ln[NO2] versus t (part (b) in Figure 5.6.1) shows us that the
reaction is not first order in NO2 because a first-order reaction would give a straight line. Having eliminated zeroth-order and first-
order behavior, we construct a plot of 1/[NO2] versus t (part (c) in Figure 5.6.1). This plot is a straight line, indicating that the
reaction is second order in NO2.

Figure 5.6.1 : The Decomposition of NO2. These plots show the decomposition of a sample of NO2 at 330°C as (a) the
concentration of NO2 versus t, (b) the natural logarithm of [NO2] versus t, and (c) 1/[NO2] versus t.
We have just determined the reaction order using data from a single experiment by plotting the concentration of the reactant as a
function of time. Because of the characteristic shapes of the lines shown in Figure 5.6.2, the graphs can be used to determine the
reaction order of an unknown reaction. In contrast, the method of initial rates required multiple experiments at different NO2
concentrations as well as accurate initial rates of reaction, which can be difficult to obtain for rapid reactions.

5.6.1 https://chem.libretexts.org/@go/page/82029
Figure 5.6.2 : Properties of Reactions That Obey Zeroth-, First-, and Second-Order Rate Laws

 Example 5.6.1

Dinitrogen pentoxide (N2O5) decomposes to NO2 and O2 at relatively low temperatures in the following reaction:

2 N2 O5 (soln) → 4N O2 (soln) + O2 (g)

This reaction is carried out in a CCl4 solution at 45°C. The concentrations of N2O5 as a function of time are listed in the
following table, together with the natural logarithms and reciprocal N2O5 concentrations. Plot a graph of the concentration
versus t, ln concentration versus t, and 1/concentration versus t and then determine the rate law and calculate the rate constant.

Time (s) [N2O5] (M) ln[N2O5] 1/[N2O5] (M−1)

0 0.0365 −3.310 27.4

600 0.0274 −3.597 36.5

1200 0.0206 −3.882 48.5

1800 0.0157 −4.154 63.7

2400 0.0117 −4.448 85.5

3000 0.00860 −4.756 116

5.6.2 https://chem.libretexts.org/@go/page/82029
Time (s) [N2O5] (M) ln[N2O5] 1/[N2O5] (M−1)

3600 0.00640 −5.051 156

Given: balanced chemical equation, reaction times, and concentrations


Asked for: graph of data, rate law, and rate constant
Strategy:
A Use the data in the table to separately plot concentration, the natural logarithm of the concentration, and the reciprocal of the
concentration (the vertical axis) versus time (the horizontal axis). Compare the graphs with those in Figure 5.6.1 to determine
the reaction order.
B Write the rate law for the reaction. Using the appropriate data from the table and the linear graph corresponding to the rate
law for the reaction, calculate the slope of the plotted line to obtain the rate constant for the reaction.

Solution
A Here are plots of [N2O5] versus t, ln[N2O5] versus t, and 1/[N2O5] versus t:

The plot of ln[N2O5] versus t gives a straight line, whereas the plots of [N2O5] versus t and 1/[N2O5] versus t do not. This means
that the decomposition of N2O5 is first order in [N2O5].
B The rate law for the reaction is therefore
rate = k[N2O5]
Calculating the rate constant is straightforward because we know that the slope of the plot of ln[A] versus t for a first-order reaction
is −k. We can calculate the slope using any two points that lie on the line in the plot of ln[N2O5] versus t. Using the points for t = 0
and 3000 s,
ln[ N2 O5 ]3000 − ln[ N2 O5 ]0
slope =
3000 s − 0 s

(−4.756) − (−3.310)
=
3000 s

−4 −1
= −4.820 × 10  s

Thus k = 4.820 × 10 −4 −1
s .

 Exercise 5.6.1
1,3-Butadiene (CH2=CH—CH=CH2; C4H6) is a volatile and reactive organic molecule used in the production of rubber. Above
room temperature, it reacts slowly to form products. Concentrations of C4H6 as a function of time at 326°C are listed in the
following table along with ln[C4H6] and the reciprocal concentrations. Graph the data as concentration versus t, ln
concentration versus t, and 1/concentration versus t. Then determine the reaction order in C4H6, the rate law, and the rate
constant for the reaction.

5.6.3 https://chem.libretexts.org/@go/page/82029
Time (s) [C4H6] (M) ln[C4H6] 1/[C4H6] (M−1)

0 1.72 × 10−2 −4.063 58.1

900 1.43 × 10−2 −4.247 69.9

1800 1.23 × 10−2 −4.398 81.3

3600 9.52 × 10−3 −4.654 105

6000 7.30 × 10−3 −4.920 137

Answer

second order in C4H6; rate = k[C4H6]2; k = 1.3 × 10−2 M−1·s−1

Summary
For a zeroth-order reaction, a plot of the concentration of any reactant versus time is a straight line with a slope of −k. For a first-
order reaction, a plot of the natural logarithm of the concentration of a reactant versus time is a straight line with a slope of −k. For
a second-order reaction, a plot of the inverse of the concentration of a reactant versus time is a straight line with a slope of k.

This page titled 5.6: Using Graphs to Determine (Integrated) Rate Laws is shared under a CC BY-NC-SA 3.0 license and was authored, remixed,
and/or curated by Anonymous.
5.7: Using Graphs to Determine Integrated Rate Laws by Anonymous is licensed CC BY-NC-SA 4.0.

5.6.4 https://chem.libretexts.org/@go/page/82029
5.7: Collision Theory
Learning Objectives
Molecules must collide in order to react.
In order to effectively initiate a reaction, collisions must be sufficiently energetic (kinetic energy) to break chemical bonds;
this energy is known as the activation energy.
As the temperature rises, molecules move faster and collide more vigorously, greatly increasing the likelihood of bond
breakage upon collision.

Collision theory explains why different reactions occur at different rates, and suggests ways to change the rate of a reaction.
Collision theory states that for a chemical reaction to occur, the reacting particles must collide with one another. The rate of the
reaction depends on the frequency of collisions. The theory also tells us that reacting particles often collide without reacting. For
collisions to be successful, reacting particles must (1) collide with (2) sufficient energy, and (3) with the proper orientation.

Requirement 1: Molecules Must Collide to React


Collision Theory provides a qualitative explanation of chemical reactions and the rates at which they occur. A basic principal of
collision theory is that, in order to react, molecules must collide. This fundamental rule guides any analysis of an ordinary reaction
mechanism. Consider a simple bimolecular step:
A + B → P roducts (5.7.1)

If the two molecules A and B are to react, they must approach closely enough to disrupt some of their existing bonds and to permit
the creation of any new ones that are needed in the products. Such an encounter is called a collision. The frequency of collisions
between A and B in a gas is proportional to the concentration of each; if [A] is doubled, the frequency of A − B collisions will
double, and doubling [B] will have the same effect. If all collisions lead to products, then the rate of a bimolecular process is first-
order in A and in B, or second-order overall:
rate = k[A][B] (5.7.2)

The need for collisions fundamental to any analysis of an ordinary reaction mechanism and explains why termolecular processes
(three species colliding and reaction) are so uncommon. The kinetic theory of gases states that for every 1000 binary collisions,
there will be only one event in which three molecules simultaneously come together. Four-way collisions are so improbable that
this process has never been demonstrated in an elementary reaction.

Illustration of the dependence of molecular collisions frequency with concentration. (Public Domain via Wikipedia).

The frequency of collisions between A and B in a gas is proportional to the concentration


of each.
Consider the reaction in the Haber process for making ammonia:
N (g) + 3 H (g) ⇌ 2 NH (g) (5.7.3)
2 2 3

the collision theory says that H and N will only react when they collide. Hence, the more frequently they collide, the faster the
2 2

rate of reaction. This can be achieved easily by either increasing the pressure on the gasses to bring H and N closer together on
2 2

average or by increasing the temperature to makes molecules move faster.

Car damage can be very expensive, especially if the person hitting your car does not have insurance. Many people have had the
experience of backing up while parallel parking and hearing that "bump". Fortunately, there is often no damage because the

5.7.1 https://chem.libretexts.org/@go/page/82048
cars were not going fast enough. But every once in a while there is a rearrangement of the body parts of a car when it is hit
with sufficient speed. Then things need to be fixed - this akin to a reaction, albeit on a macroscopic scale.

Energy for thermally activated reactions come from collisions of species.

Requirement 2: Not all Collisions are Sufficiently Energetic


In the Haber process (Equation 5.7.3) at 300 K only 1 in 10 collisions between H and N results in a reaction! At 800 K, this
11
2 2

increases to 1 in 10 collisions resulting in a reaction. Hence, while the collisions are needed for a reaction, other aspects
4

contribute. Reacting particles can form products when they collide with one another provided those collisions have enough kinetic
energy and the correct orientation. Particles that lack the necessary kinetic energy may collide, but the particles will simply bounce
off one another unchanged. Figure 5.7.1 illustrates the difference. In the first collision, the particles bounce off one another and no
rearrangement of atoms has occurred. The second collision occurs with greater kinetic energy, and so the bond between the two red
atoms breaks. One red atom bonds with the other molecule as one product, while the single red atom is the other product. The first
collision is called an ineffective collision, while the second collision is called an effective collision.

Figure 5.7.1 : An ineffective collision (A) is one that does not result in product formation. An effective collision (B) is one in which
chemical bonds are broken and a product is formed.
For a gas at room temperature and normal atmospheric pressure, there are about 1033 collisions in each cubic centimeter of space
every second. If every collision between two reactant molecules yielded products, all reactions would be complete in a fraction of a
second. For example, when two billiard balls collide, they simply bounce off of each other. This is the most likely outcome if the
reaction between A and B requires a significant disruption or rearrangement of the bonds between their atoms. In order to
effectively initiate a reaction, collisions must be sufficiently energetic (or have sufficient kinetic energy) to bring about this bond
disruption.
A reaction will not take place unless the particles collide with a certain minimum energy called the activation energy of the
reaction. Activation energy is the minimum energy required to make a reaction occur. This can be illustrated on an energy profile
for the reaction. An energy profile for a simple exothermic reaction is given in the Figure 5.7.2.

Figure 5.7.2 : Exothermic reaction profile

5.7.2 https://chem.libretexts.org/@go/page/82048
If the particles collide with less energy than the activation energy, nothing interesting happens. They bounce apart. The activation
energy can be thought of as a barrier to the reaction. Only those collisions with energies equal to or greater than the activation
energy result in a reaction.

Any chemical reaction results in the breaking of some bonds (which requires energy) and
the formation of new ones (which releases energy). Some bonds must be broken before
new ones can be formed. Activation energy is involved in breaking some of the original
bonds. If a collision is relatively gentle, there is insufficient energy available to initiate the
bond-breaking process, and thus the particles do not react.
Energetic collisions between molecules cause interatomic bonds to stretch and bend, temporarily weakening them so that they
become more susceptible to cleavage. Distortion of the bonds can expose their associated electron clouds to interactions with other
reactants that might lead to the formation of new bonds.

Figure 5.7.3 : Anatomy of a collision.


Chemical bonds have some of the properties of mechanical springs: their potential energies depend on the extent to which they are
stretched or compressed. Each atom-to-atom bond can be described by a potential energy diagram that shows how its energy
changes with its length. When the bond absorbs energy (either from heating or through a collision), it is elevated to a higher
quantized vibrational state (indicated by the horizontal lines) that weakens the bond as its length oscillates between the extended
limits corresponding to the curve in Figure 5.7.3.

When the bond absorbs energy (either from heating or through a collision), it is elevated
to a higher quantized vibrational state (indicated by the horizontal lines) that weakens the
bond.
A particular collision will typically excite a number of bonds in this way. Within about 10–13 seconds, this excitation is distributed
among the other bonds in the molecule in complex and unpredictable ways that can concentrate the added energy at a particularly
vulnerable point. The affected bond can stretch and bend farther, making it more susceptible to cleavage. Even if the bond does not
break by pure stretching, it can become distorted or twisted so as to expose nearby electron clouds to interactions with other
reactants that might encourage a reaction.

Example 5.7.1

a. Draw a simple energy profile for an exothermic reaction in which 100 kJ mol-1 is evolved, and which has an activation
energy of 50 kJ mol-1 .
b. Draw a simple energy profile for an endothermic reaction in which 50 kJ mol-1 is absorbed and which has an activation
energy of 100 kJ mol-1
c. Explain why all reactions have an activation energy, using your knowledge of collision theory.

Answer
c: All reactions have an activation energy because energy is required to make the reactants combine in a way that will cause
the reaction. No chemical process can take place without having at least a little energy to get things started.

5.7.3 https://chem.libretexts.org/@go/page/82048
Exercise 5.7.1

To increase the rate of a reaction, there must be (select one):


a. Decrease in the frequency of collisions
b. An Increase in the frequency of collisions.
c. A decrease in the frequency of effective collisions
d. An increase in the frequency of effective collisions

Answer
D

The Maxwell-Boltzmann Distribution


Because of the key role of activation energy in deciding whether a collision will result in a reaction, it is useful to know the
proportion of the particles present with high enough energies to react when they collide. In any system, the particles present
will have a very wide range of energies. For gases, this can be shown on a graph called the Maxwell-Boltzmann distribution, a
plot showing the number of particles with each particular energy.

The area under the curve measures of the total number of particles present. Remember that for a reaction to occur, particles
must collide with energies equal to or greater than the activation energy for the reaction. The activation energy is marked on
the Maxwell-Boltzmann distribution with a green line:

Notice that the large majority of the particles have insufficient energy to react when they collide. To enable them to react,
either the shape of the curve must be altered, or the activation energy shifted further to the left to lower energies.

Requirement 3: Not all Collisions are Sufficiently Oriented


Even if two molecules collide with sufficient activation energy, there is no guarantee that the collision will be successful. In fact,
the collision theory says that not every collision is successful, even if molecules are moving with enough energy. The reason for
this is because molecules also need to collide with the right orientation, so that the proper atoms line up with one another, and
bonds can break and re-form in the necessary fashion. However, because molecules in the liquid and gas phase are in constant,
random motion, there is always the probability that two molecules will collide in just the right way for them to react.

5.7.4 https://chem.libretexts.org/@go/page/82048
Consider a simple reaction involving a collision between two molecules: for example, ethene, CH 2
=CH
2
, and hydrogen
chloride, H C l. These react to give chloroethane as shown:
H C=CH + HCl → CH CH Cl (5.7.4)
2 2 3 2

As a result of the collision between the two molecules, the double bond in ethene is converted into a single bond. A hydrogen atom
is now attached to one of the carbons and a chlorine atom to the other. The reaction can only happen if the hydrogen end of the H-
Cl bond approaches the carbon-carbon double bond. No other collision between the two molecules produces the same effect. The
two simply bounce off each other.

Figure 5.7.4 : Of the collisions shown in the diagram, only collision 1 may possibly lead on to a reaction (if enough kinetic energy
is involved). The other three collisions will not lead to a reaction irrespective of the kinetic energy involved.
With no knowledge of the reaction mechanism, one might wonder why collision 2 would be unsuccessful. The double bond has a
high concentration of negative charge around it due to the electrons in the bonds. The approaching chlorine atom is also partially
negative due to dipole created by the electronegativity difference between it and hydrogen. The repulsion simply causes the
molecules to bounce off each other. In any collision involving unsymmetrical species, the way they hit each other is important in
determining whether a reaction occurs.

Unimolecular processes also begin with a Collision


Until about 1921, chemists did not understand the role of collisions in unimolecular processes. It turns out that the mechanisms
of such reactions are actually quite complicated, and that at very low pressures they do follow second-order kinetics. Such
reactions are more properly described as pseudounimolecular. The cyclopropane isomerization described in Example 1 is
typical of many decomposition reactions found to follow first-order kinetics, implying that the process is unimolecular.
Consider, for example, the isomerization of cyclopropane to propene, which takes place at fairly high temperatures in the gas
phase:

The collision-to-product sequence can be conceptualized in the following [grossly oversimplified] way:

Note that
For simplicity, the hydrogen atoms are not shown here. This is reasonable because C–C bonds are weaker then C–H bonds,
which are less likely to be affected.
The collision at 1 is most likely with another cyclopropane molecule, but because no part of the colliding molecule gets
incorporated into the product, it can in principle be a noble gas or some other non-reacting species;
Although the C–C bonds in cyclopropane are all identical, the instantaneous localization of the collisional energy can
distort the molecule in various ways (2), leading to a configuration sufficiently unstable to initiate the rearrangement to the
product.

Of course, the more critical this orientational requirement is, like it is for larger or more complex molecules, the fewer collisions
there will be that will be effective. An effective collision is defined as one in which molecules collide with sufficient energy and
proper orientation, so that a reaction occurs.

5.7.5 https://chem.libretexts.org/@go/page/82048
Exercise 5.7.2

According to collision theory, what three criteria are needed to be met before a bimolecular reaction can take place?
a. The orientation probability factor must be 1.
b. The collision energy must be greater than the activation energy for the reaction.
c. The collision must occur in the proper orientation.
d. The collision frequency must be greater than the frequency factor for the reaction.
e. A collision between the reactants must occur.

Exercise 5.7.3

How will the given change affect the rate of an elementary reaction.
a. An increase in temperature.
b. An increase in the activation energy for the reaction
c. An increase in the reactant concentration.
d. An increase in the size of one of the reactants in a bimolecular reaction

Summary
Collision Theory provides a qualitative explanation of chemical reactions and the rates at which they occur. For a chemical reaction
to occur, an energy threshold must be overcome, and the reacting species must also have the correct spatial orientation. Factors that
increase the rate of a reaction must influence at least one of the following:
How often collisions occur - more frequent collisions will mean a faster rate.
More effective collisions in terms of collisions occurring with sufficient energy
More effective collisions in terms of collisions occurring with the proper orientation.
Two more topics must be examined before these can be discussed in depth: reaction mechanisms and the concept of threshold
energy.

Contributors and Attributions


Peggy Lawson (Oxbow Prairie Heights School)
Stephen Lower, Professor Emeritus (Simon Fraser U.) Chem1 Virtual Textbook

5.7: Collision Theory is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

5.7.6 https://chem.libretexts.org/@go/page/82048
5.8: Temperature and Rate
 Learning Objectives
To understand why and how chemical reactions occur.

It is possible to use kinetics studies of a chemical system, such as the effect of changes in reactant concentrations, to deduce events
that occur on a microscopic scale, such as collisions between individual particles. Such studies have led to the collision model of
chemical kinetics, which is a useful tool for understanding the behavior of reacting chemical species. The collision model explains
why chemical reactions often occur more rapidly at higher temperatures. For example, the reaction rates of many reactions that
occur at room temperature approximately double with a temperature increase of only 10°C. In this section, we will use the collision
model to analyze this relationship between temperature and reaction rates. Before delving into the relationship between temperature
and reaction rate, we must discuss three microscopic factors that influence the observed macroscopic reaction rates.

Microscopic Factor 1: Collisional Frequency


Central to collision model is that a chemical reaction can occur only when the reactant molecules, atoms, or ions collide. Hence, the
observed rate is influence by the frequency of collisions between the reactants. The collisional frequency is the average rate in
which two reactants collide for a given system and is used to express the average number of collisions per unit of time in a defined
system. While deriving the collisional frequency (Z ) between two species in a gas is straightforward, it is beyond the scope of
AB

this text and the equation for collisional frequency of A and B is the following:
−−−−−−
2
8π kB T
ZAB = NA NB (rA + rB ) √ (5.8.1)
μAB

with
NA and N are the numbers of A and B molecules in the system, respectively
B

ra and r are the radii of molecule A and B , respectively


b

kB is the Boltzmann constant k = 1.380 x 10-23 Joules Kelvin


B

T is the temperature in Kelvin


mA mB
μAB is calculated via μ =
AB mA +mB

The specifics of Equation 5.8.1 are not important for this conversation, but it is important to identify that Z increases with
AB

increasing density (i.e., increasing N and N ), with increasing reactant size (r and r ), with increasing velocities (predicted via
A B a b

Kinetic Molecular Theory), and with increasing temperature (although weakly because of the square root function).

I'm not a robot


reCAPTCHA
Privacy - Terms

About this page

Our systems have detected unusual traffic from your


computer network This page checks to see if it's really you
A Video Discussing Collision Theory of Kinetics: Collusion Theory of Kinetics (opens in new window) [youtu.be]

5.8.1 https://chem.libretexts.org/@go/page/81884
Microscopic Factor 2: Activation Energy
Previously, we discussed the kinetic molecular theory of gases, which showed that the average kinetic energy of the particles of a
gas increases with increasing temperature. Because the speed of a particle is proportional to the square root of its kinetic energy,
increasing the temperature will also increase the number of collisions between molecules per unit time. What the kinetic molecular
theory of gases does not explain is why the reaction rate of most reactions approximately doubles with a 10°C temperature
increase. This result is surprisingly large considering that a 10°C increase in the temperature of a gas from 300 K to 310 K
increases the kinetic energy of the particles by only about 4%, leading to an increase in molecular speed of only about 2% and a
correspondingly small increase in the number of bimolecular collisions per unit time.
The collision model of chemical kinetics explains this behavior by introducing the concept of activation energy (E ). We will a

define this concept using the reaction of NO with ozone, which plays an important role in the depletion of ozone in the ozone
layer:

NO(g) + O (g) → NO (g) + O (g)


3 2 2

Increasing the temperature from 200 K to 350 K causes the rate constant for this particular reaction to increase by a factor of more
than 10, whereas the increase in the frequency of bimolecular collisions over this temperature range is only 30%. Thus something
other than an increase in the collision rate must be affecting the reaction rate.
Experimental rate law for this reaction is

rate = k[NO][ O ]
3

and is used to identify how the reaction rate (not the rate constant) vares with concentration. The rate constant, however, does vary
with temperature. Figure 5.8.1 shows a plot of the rate constant of the reaction of NO with O at various temperatures. The
3

relationship is not linear but instead resembles the relationships seen in graphs of vapor pressure versus temperature (e.g, the
Clausius-Claperyon equation). In all three cases, the shape of the plots results from a distribution of kinetic energy over a
population of particles (electrons in the case of conductivity; molecules in the case of vapor pressure; and molecules, atoms, or ions
in the case of reaction rates). Only a fraction of the particles have sufficient energy to overcome an energy barrier.

Figure 5.8.1 : Rate Constant versus Temperature for the Reaction of NO with O The nonlinear shape of the curve is caused by a
3

distribution of kinetic energy over a population of molecules. Only a fraction of the particles have enough energy to overcome an
energy barrier, but as the temperature is increased, the size of that fraction increases. (CC BY-SA-NC; anonymous)
In the case of vapor pressure, particles must overcome an energy barrier to escape from the liquid phase to the gas phase. This
barrier corresponds to the energy of the intermolecular forces that hold the molecules together in the liquid. In conductivity, the
barrier is the energy gap between the filled and empty bands. In chemical reactions, the energy barrier corresponds to the amount of
energy the particles must have to react when they collide. This energy threshold, called the activation energy, was first postulated
in 1888 by the Swedish chemist Svante Arrhenius (1859–1927; Nobel Prize in Chemistry 1903). It is the minimum amount of
energy needed for a reaction to occur. Reacting molecules must have enough energy to overcome electrostatic repulsion, and a
minimum amount of energy is required to break chemical bonds so that new ones may be formed. Molecules that collide with less
than the threshold energy bounce off one another chemically unchanged, with only their direction of travel and their speed altered
by the collision. Molecules that are able to overcome the energy barrier are able to react and form an arrangement of atoms called
the activated complex or the transition state of the reaction. The activated complex is not a reaction intermediate; it does not last
long enough to be detected readily.

5.8.2 https://chem.libretexts.org/@go/page/81884
Any phenomenon that depends on the distribution of thermal energy in a population of
particles has a nonlinear temperature dependence.
We can graph the energy of a reaction by plotting the potential energy of the system as the reaction progresses. Figure 5.8.2 shows
a plot for the NO–O3 system, in which the vertical axis is potential energy and the horizontal axis is the reaction coordinate, which
indicates the progress of the reaction with time. The activated complex is shown in brackets with an asterisk. The overall change in
potential energy for the reaction (ΔE) is negative, which means that the reaction releases energy. (In this case, ΔE is −200.8
kJ/mol.) To react, however, the molecules must overcome the energy barrier to reaction (E is 9.6 kJ/mol). That is, 9.6 kJ/mol must
a

be put into the system as the activation energy. Below this threshold, the particles do not have enough energy for the reaction to
occur.

Figure 5.8.2 : Energy of the Activated Complex for the NO–O3 System. The diagram shows how the energy of this system varies as
the reaction proceeds from reactants to products. Note the initial increase in energy required to form the activated complex. (CC
BY-SA-NC; anonymous)
Figure 5.8.3a illustrates the general situation in which the products have a lower potential energy than the reactants. In contrast,
Figure 5.8.3b illustrates the case in which the products have a higher potential energy than the reactants, so the overall reaction
requires an input of energy; that is, it is energetically uphill, and \(ΔE > 0\). Although the energy changes that result from a reaction
can be positive, negative, or even zero, in most cases an energy barrier must be overcome before a reaction can occur. This means
that the activation energy is almost always positive; there is a class of reactions called barrierless reactions, but those are discussed
elsewhere.

Figure 5.8.3 : Differentiating between E and ΔE. The potential energy diagrams for a reaction with (a) ΔE < 0 and (b) ΔE > 0
a

illustrate the change in the potential energy of the system as reactants are converted to products. In both cases, E is positive. For a
a

reaction such as the one shown in (b), Ea must be greater than ΔE. (CC BY-SA-NC; anonymous)

For similar reactions under comparable conditions, the one with the smallest Ea will
occur most rapidly.
Whereas ΔE is related to the tendency of a reaction to occur spontaneously, E gives us information about the reaction rate and
a

how rapidly the reaction rate changes with temperature. For two similar reactions under comparable conditions, the reaction with

5.8.3 https://chem.libretexts.org/@go/page/81884
the smallest E will occur more rapidly.
a

Figure 5.8.4 shows both the kinetic energy distributions and a potential energy diagram for a reaction. The shaded areas show that
at the lower temperature (300 K), only a small fraction of molecules collide with kinetic energy greater than Ea; however, at the
higher temperature (500 K) a much larger fraction of molecules collide with kinetic energy greater than Ea. Consequently, the
reaction rate is much slower at the lower temperature because only a relatively few molecules collide with enough energy to
overcome the potential energy barrier.

Figure 5.8.4 : Surmounting the Energy Barrier to a Reaction. This chart juxtaposes the energy distributions of lower-temperature
(300 K) and higher-temperature (500 K) samples of a gas against the potential energy diagram for a reaction. Only those molecules
in the shaded region of the energy distribution curve have E > E and are therefore able to cross the energy barrier separating
a

reactants and products. The fraction of molecules with E > E is much greater at 500 K than at 300 K, so the reaction will occur
a

much more rapidly at 500 K. (CC BY-SA-NC; anonymous)


Energy is on the y axis while reaction coordinate and fraction of molecules with a particular kinetic energy E are on the x axis.

I'm not a robot


reCAPTCHA
Privacy - Terms

About this page

Our systems have detected unusual traffic from your


computer network This page checks to see if it's really you
Video Discussing Transition State Theory: Transition State Theory(opens in new window) [youtu.be]

Microscopic Factor 3: Sterics


Even when the energy of collisions between two reactant species is greater than E , most collisions do not produce a reaction. The
a

probability of a reaction occurring depends not only on the collision energy but also on the spatial orientation of the molecules
when they collide. For NO and O to produce NO and O , a terminal oxygen atom of O must collide with the nitrogen atom of
3 2 2 3

NO at an angle that allows O to transfer an oxygen atom to NO to produce NO (Figure 5.8.4). All other collisions produce no
3 2

reaction. Because fewer than 1% of all possible orientations of NO and O result in a reaction at kinetic energies greater than E ,
3 a

most collisions of NO and O are unproductive. The fraction of orientations that result in a reaction is called the steric factor (ρ)
3

and its value can range from ρ = 0 (no orientations of molecules result in reaction) to ρ = 1 (all orientations result in reaction).

5.8.4 https://chem.libretexts.org/@go/page/81884
Figure 5.8.4 : The Effect of Molecular Orientation on the Reaction of NO and O . Most collisions of NO and O molecules occur
3 3

with an incorrect orientation for a reaction to occur. Only those collisions in which the N atom of NO collides with one of the
terminal O atoms of O are likely to produce NO and O , even if the molecules collide with E > E . (CC BY-SA-NC;
3 2 2 a

anonymous)

Macroscopic Behavior: The Arrhenius Equation


The collision model explains why most collisions between molecules do not result in a chemical reaction. For example, nitrogen
and oxygen molecules in a single liter of air at room temperature and 1 atm of pressure collide about 1030 times per second. If
every collision produced two molecules of NO, the atmosphere would have been converted to NO and then NO a long time ago. 2

Instead, in most collisions, the molecules simply bounce off one another without reacting, much as marbles bounce off each other
when they collide.
For an A + B elementary reaction, all three microscopic factors discussed above that affect the reaction rate can be summarized in
a single relationship:

rate = (collision frequency) × (steric factor) × (fraction of collisions with E > Ea )

where
rate = k[A][B] (5.8.2)

Arrhenius used these relationships to arrive at an equation that relates the magnitude of the rate constant for a reaction to the
temperature, the activation energy, and the constant, A , called the frequency factor:
−Ea /RT
k = Ae (5.8.3)

The frequency factor is used to convert concentrations to collisions per second (scaled by the steric factor). Because the frequency
of collisions depends on the temperature, A is actually not constant (Equation 5.8.1). Instead, A increases slightly with
temperature as the increased kinetic energy of molecules at higher temperatures causes them to move slightly faster and thus
undergo more collisions per unit time.
Equation 5.8.3 is known as the Arrhenius equation and summarizes the collision model of chemical kinetics, where T is the
absolute temperature (in K) and R is the ideal gas constant [8.314 J/(K·mol)]. E indicates the sensitivity of the reaction to changes
a

in temperature. The reaction rate with a large E increases rapidly with increasing temperature, whereas the reaction rate with a
a

smaller E increases much more slowly with increasing temperature.


a

If we know the reaction rate at various temperatures, we can use the Arrhenius equation to calculate the activation energy. Taking
the natural logarithm of both sides of Equation 5.8.3,
Ea
ln k = ln A + (− ) (5.8.4)
RT

Ea 1
= ln A + [(− )( )] (5.8.5)
R T

Equation 5.8.5 is the equation of a straight line,

y = mx + b

5.8.5 https://chem.libretexts.org/@go/page/81884
where y = ln k and x = 1/T . This means that a plot of ln k versus 1/T is a straight line with a slope of −E a /R and an intercept
of ln A . In fact, we need to measure the reaction rate at only two temperatures to estimate E . a

Knowing the E at one temperature allows us to predict the reaction rate at other temperatures. This is important in cooking and
a

food preservation, for example, as well as in controlling industrial reactions to prevent potential disasters. The procedure for
determining E from reaction rates measured at several temperatures is illustrated in Example 5.8.1.
a

I'm not a robot


reCAPTCHA
Privacy - Terms

About this page

Our systems have detected unusual traffic from your


computer network This page checks to see if it's really you
A Video Discussing The Arrhenius Equation: The Arrhenius Equation(opens in new window) [youtu.be]

 Example 5.8.1: Chirping Tree Crickets

Many people believe that the rate of a tree cricket’s chirping is related to temperature. To see whether this is true, biologists
have carried out accurate measurements of the rate of tree cricket chirping (f ) as a function of temperature (T ). Use the data in
the following table, along with the graph of ln[chirping rate] versus 1/T to calculate E for the biochemical reaction that
a

controls cricket chirping. Then predict the chirping rate on a very hot evening, when the temperature is 308 K (35°C, or 95°F).
Chirping Tree Crickets Frequency Table
Frequency (f; chirps/min) ln f T (K) 1/T (K)

200 5.30 299 3.34 × 10−3

179 5.19 298 3.36 × 10−3

158 5.06 296 3.38 × 10−3

141 4.95 294 3.40 × 10−3

126 4.84 293 3.41 × 10−3

112 4.72 292 3.42 × 10−3

100 4.61 290 3.45 × 10−3

89 4.49 289 3.46 × 10−3

79 4.37 287 3.48 × 10−3

Given: chirping rate at various temperatures


Asked for: activation energy and chirping rate at specified temperature

Strategy:
A. From the plot of ln f versus 1/T , calculate the slope of the line (−Ea/R) and then solve for the activation energy.
B. Express Equation 5.8.5 in terms of k1 and T1 and then in terms of k2 and T2.
C. Subtract the two equations; rearrange the result to describe k2/k1 in terms of T2 and T1.

5.8.6 https://chem.libretexts.org/@go/page/81884
D. Using measured data from the table, solve the equation to obtain the ratio k2/k1. Using the value listed in the table for k1,
solve for k2.

Solution
A If cricket chirping is controlled by a reaction that obeys the Arrhenius equation, then a plot of ln f versus 1/T should give a
straight line (Figure 5.8.6).

Figure 5.8.6 : Graphical Determination of E for Tree Cricket Chirping. When the natural logarithm of the rate of tree cricket
a

chirping is plotted versus 1/T, a straight line results. The slope of the line suggests that the chirping rate is controlled by a
single reaction with an E of 55 kJ/mol. (CC BY-SA-NC; anonymous)
a

Also, the slope of the plot of ln f versus 1/T should be equal to −Ea /R . We can use the two endpoints in Figure 5.8.6 to
estimate the slope:
Δ ln f
slope =
Δ(1/T )

5.30 − 4.37
=
−3 −1 −3 −1
3.34 × 10  K − 3.48 × 10  K

0.93
=
−3 −1
−0.14 × 10  K

3
= −6.6 × 10  K

A computer best-fit line through all the points has a slope of −6.67 × 103 K, so our estimate is very close. We now use it to
solve for the activation energy:
Ea = −(slope)(R)

8.314 J 1 KJ
3
= −(−6.6 × 10  K) ( )( )
K ⋅ mol 1000 J

55 kJ
=
mol

B If the activation energy of a reaction and the rate constant at one temperature are known, then we can calculate the reaction
rate at any other temperature. We can use Equation 5.8.5 to express the known rate constant (k ) at the first temperature (T )
1 1

as follows:
Ea
ln k1 = ln A −
RT1

Similarly, we can express the unknown rate constant (k ) at the second temperature (T ) as follows:
2 2

Ea
ln k2 = ln A −
RT2

5.8.7 https://chem.libretexts.org/@go/page/81884
C These two equations contain four known quantities (Ea, T1, T2, and k1) and two unknowns (A and k2). We can eliminate A by
subtracting the first equation from the second:
Ea Ea
ln k2 − ln k1 = (ln A − ) − (ln A − )
RT2 RT1

Ea Ea
=− +
RT2 RT1

Then
k2 Ea 1 1
ln = ( − )
k1 R T1 T2

D To obtain the best prediction of chirping rate at 308 K (T2), we try to choose for T1 and k1 the measured rate constant and
corresponding temperature in the data table that is closest to the best-fit line in the graph. Choosing data for T1 = 296 K, where
f = 158, and using the E calculated previously,
a

kT Ea 1 1
2
ln = ( − )
kT R T1 T2
1

55 kJ/mol 1000 J 1 1
= ( )( − )
8.314 J/(K ⋅ mol) 1 kJ 296 K 308 K

= 0.87

Thus k308/k296 = 2.4 and k308 = (2.4)(158) = 380, and the chirping rate on a night when the temperature is 308 K is predicted to
be 380 chirps per minute.

 Exercise 5.8.1A

The equation for the decomposition of NO to NO and O is second order in NO :


2 2 2

2 NO (g) → 2 NO(g) + O (g)


2 2

Data for the reaction rate as a function of temperature are listed in the following table. Calculate Ea for the reaction and the
rate constant at 700 K.
Data for the reaction rate as a function of temperature
T (K) k (M−1·s−1)

592 522

603 755

627 1700

652 4020

656 5030

Answer
Ea = 114 kJ/mol; k700= 18,600 M−1·s−1 = 1.86 × 104 M−1·s−1.

 Exercise 5.8.1B
What E results in a doubling of the reaction rate with a 10°C increase in temperature from 20° to 30°C?
a

Answer
about 51 kJ/mol

5.8.8 https://chem.libretexts.org/@go/page/81884
I'm not a robot
reCAPTCHA
Privacy - Terms

About this page

Our systems have detected unusual traffic from your


computer network This page checks to see if it's really you
A Video Discussing Graphing Using the Arrhenius Equation: Graphing Using the Arrhenius Equation (opens in new window)
[youtu.be] (opens in new window)

Summary
For a chemical reaction to occur, an energy threshold must be overcome, and the reacting species must also have the correct spatial
orientation. The Arrhenius equation is k = Ae −Ea /RT
. A minimum energy (activation energy,vE ) is required for a collision
a

between molecules to result in a chemical reaction. Plots of potential energy for a system versus the reaction coordinate show an
energy barrier that must be overcome for the reaction to occur. The arrangement of atoms at the highest point of this barrier is the
activated complex, or transition state, of the reaction. At a given temperature, the higher the Ea, the slower the reaction. The
fraction of orientations that result in a reaction is the steric factor. The frequency factor, steric factor, and activation energy are
related to the rate constant in the Arrhenius equation: k = Ae −Ea /RT
. A plot of the natural logarithm of k versus 1/T is a straight
line with a slope of −Ea/R.

5.8: Temperature and Rate is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
14.5: Temperature and Rate is licensed CC BY-NC-SA 3.0.

5.8.9 https://chem.libretexts.org/@go/page/81884
5.9: Reaction Mechanisms
 Learning Objectives
To determine the individual steps of a simple reaction.

One of the major reasons for studying chemical kinetics is to use measurements of the macroscopic properties of a system, such as
the rate of change in the concentration of reactants or products with time, to discover the sequence of events that occur at the
molecular level during a reaction. This molecular description is the mechanism of the reaction; it describes how individual atoms,
ions, or molecules interact to form particular products. The stepwise changes are collectively called the reaction mechanism.
In an internal combustion engine, for example, isooctane reacts with oxygen to give carbon dioxide and water:

2C H (l) + 25 O (g) ⟶ 16 CO (g) + 18 H O(g) (5.9.1)


8 18 2 2 2

For this reaction to occur in a single step, 25 dioxygen molecules and 2 isooctane molecules would have to collide simultaneously
and be converted to 34 molecules of product, which is very unlikely. It is more likely that a complex series of reactions takes place
in a stepwise fashion. Each individual reaction, which is called an elementary reaction, involves one, two, or (rarely) three atoms,
molecules, or ions. The overall sequence of elementary reactions is the mechanism of the reaction. The sum of the individual steps,
or elementary reactions, in the mechanism must give the balanced chemical equation for the overall reaction.

The overall sequence of elementary reactions is the mechanism of the reaction.

Molecularity and the Rate-Determining Step


To demonstrate how the analysis of elementary reactions helps us determine the overall reaction mechanism, we will examine the
much simpler reaction of carbon monoxide with nitrogen dioxide.
\[\ce{NO2(g) + CO(g) -> NO(g) + CO2 (g)} \label{14.6.2} \]
From the balanced chemical equation, one might expect the reaction to occur via a collision of one molecule of NO with a 2

molecule of CO that results in the transfer of an oxygen atom from nitrogen to carbon. The experimentally determined rate law for
the reaction, however, is as follows:
2
rate = k[NO ] (5.9.2)
2

The fact that the reaction is second order in [NO ] and independent of [CO] tells us that it does not occur by the simple collision
2

model outlined previously. If it did, its predicted rate law would be

rate = k[ NO ][CO].
2

The following two-step mechanism is consistent with the rate law if step 1 is much slower than step 2:
two-step mechanism
slow
step 1 elementary reaction
N O2 + N O2 −−→ N O3 + NO

step 2 N O3 + CO → N O2 + CO2 elementary reaction


––––––––––––––––––––––––––

sum N O2 + CO → NO + CO2 overall reaction

According to this mechanism, the overall reaction occurs in two steps, or elementary reactions. Summing steps 1 and 2 and
canceling on both sides of the equation gives the overall balanced chemical equation for the reaction. The NO molecule is an
3

intermediate in the reaction, a species that does not appear in the balanced chemical equation for the overall reaction. It is formed
as a product of the first step but is consumed in the second step.

The sum of the elementary reactions in a reaction mechanism must give the overall
balanced chemical equation of the reaction.

5.9.1 https://chem.libretexts.org/@go/page/81885
Using Molecularity to Describe a Rate Law
The molecularity of an elementary reaction is the number of molecules that collide during that step in the mechanism. If there is
only a single reactant molecule in an elementary reaction, that step is designated as unimolecular; if there are two reactant
molecules, it is bimolecular; and if there are three reactant molecules (a relatively rare situation), it is termolecular. Elementary
reactions that involve the simultaneous collision of more than three molecules are highly improbable and have never been observed
experimentally. (To understand why, try to make three or more marbles or pool balls collide with one another simultaneously!)

Figure 5.9.1 : The Basis for Writing Rate Laws of Elementary Reactions. This diagram illustrates how the number of possible
collisions per unit time between two reactant species, A and B, depends on the number of A and B particles present. The number of
collisions between A and B particles increases as the product of the number of particles, not as the sum. This is why the rate law for
an elementary reaction depends on the product of the concentrations of the species that collide in that step. (CC BY-NC-SA;
anonymous)
Writing the rate law for an elementary reaction is straightforward because we know how many molecules must collide
simultaneously for the elementary reaction to occur; hence the order of the elementary reaction is the same as its molecularity
(Table 5.9.1). In contrast, the rate law for the reaction cannot be determined from the balanced chemical equation for the overall
reaction. The general rate law for a unimolecular elementary reaction (A → products) is

rate = k[A].

For bimolecular reactions, the reaction rate depends on the number of collisions per unit time, which is proportional to the product
of the concentrations of the reactants, as shown in Figure 5.9.1. For a bimolecular elementary reaction of the form A + B →
products, the general rate law is

rate = k[A][B].

Table 5.9.1 : Common Types of Elementary Reactions and Their Rate Laws
Elementary Reaction Molecularity Rate Law Reaction Order

A → products unimolecular rate = k[A] first

2A → products bimolecular rate = k[A]2 second

A + B → products bimolecular rate = k[A][B] second

2A + B → products termolecular rate = k[A]2[B] third

A + B + C → products termolecular rate = k[A][B][C] third

For elementary reactions, the order of the elementary reaction is the same as its
molecularity. In contrast, the rate law cannot be determined from the balanced chemical
equation for the overall reaction (unless it is a single step mechanism and is therefore
also an elementary step).

Identifying the Rate-Determining Step


Note the important difference between writing rate laws for elementary reactions and the balanced chemical equation of the overall
reaction. Because the balanced chemical equation does not necessarily reveal the individual elementary reactions by which the
reaction occurs, we cannot obtain the rate law for a reaction from the overall balanced chemical equation alone. In fact, it is the rate
law for the slowest overall reaction, which is the same as the rate law for the slowest step in the reaction mechanism, the rate-
determining step, that must give the experimentally determined rate law for the overall reaction.This statement is true if one step
is substantially slower than all the others, typically by a factor of 10 or more. If two or more slow steps have comparable rates, the
experimentally determined rate laws can become complex. Our discussion is limited to reactions in which one step can be

5.9.2 https://chem.libretexts.org/@go/page/81885
identified as being substantially slower than any other. The reason for this is that any process that occurs through a sequence of
steps can take place no faster than the slowest step in the sequence. In an automotive assembly line, for example, a component
cannot be used faster than it is produced. Similarly, blood pressure is regulated by the flow of blood through the smallest passages,
the capillaries. Because movement through capillaries constitutes the rate-determining step in blood flow, blood pressure can be
regulated by medications that cause the capillaries to contract or dilate. A chemical reaction that occurs via a series of elementary
reactions can take place no faster than the slowest step in the series of reactions.

Rate-determining step. The phenomenon of a rate-determining step can be compared to a succession of funnels. The smallest-
diameter funnel controls the rate at which the bottle is filled, whether it is the first or the last in the series. Pouring liquid into the
first funnel faster than it can drain through the smallest results in an overflow. (CC BY-NC-SA; anonymous)
Look at the rate laws for each elementary reaction in our example as well as for the overall reaction.
rate laws for each elementary reaction in our example as well as for the overall reaction.
k1
2
step 1 N O2 + N O2 −
→ N O3 + NO rate = k1 [N O2 ]  (predicted)

k2

step 2 N O3 + CO − → N O2 + CO2 rate = k2 [N O3 ][CO] (predicted)


––––––––––––––––––––––––––

k
2
sum N O2 + CO → NO + CO2 rate = k[N O2 ]  (observed)

The experimentally determined rate law for the reaction of N O with C O is the same as the predicted rate law for step 1. This tells
2

us that the first elementary reaction is the rate-determining step, so k for the overall reaction must equal k . That is, NO3 is formed
1

slowly in step 1, but once it is formed, it reacts very rapidly with CO in step 2.
Sometimes chemists are able to propose two or more mechanisms that are consistent with the available data. If a proposed
mechanism predicts the wrong experimental rate law, however, the mechanism must be incorrect.

 Example 5.9.1: A Reaction with an Intermediate

In an alternative mechanism for the reaction of NO with CO with N


2 2
O
4
appearing as an intermediate.
alternative mechanism for the reaction of NO with CO with N
2 2
O
4
appearing as an intermediate.
k1
step 1
N O2 + N O2 −
→ N2 O4

k2
step 2 N2 O4 + CO − → NO + N O2 + CO2
–––––––––––––––––––––––––––––––––

sum N O2 + CO → NO + CO2

Write the rate law for each elementary reaction. Is this mechanism consistent with the experimentally determined rate law (rate
= k[NO2]2)?
Given: elementary reactions

5.9.3 https://chem.libretexts.org/@go/page/81885
Asked for: rate law for each elementary reaction and overall rate law

Strategy:
A. Determine the rate law for each elementary reaction in the reaction.
B. Determine which rate law corresponds to the experimentally determined rate law for the reaction. This rate law is the one
for the rate-determining step.

Solution
A The rate law for step 1 is rate = k1[NO2]2; for step 2, it is rate = k2[N2O4][CO].
B If step 1 is slow (and therefore the rate-determining step), then the overall rate law for the reaction will be the same: rate =
k1[NO2]2. This is the same as the experimentally determined rate law. Hence this mechanism, with N2O4 as an intermediate,
and the one described previously, with NO3 as an intermediate, are kinetically indistinguishable. In this case, further
experiments are needed to distinguish between them. For example, the researcher could try to detect the proposed
intermediates, NO3 and N2O4, directly.

 Exercise 5.9.1

Iodine monochloride (ICl) reacts with H as follows:


2

2 ICl(l) + H (g) → 2 HCl(g) + I (s)


2 2

The experimentally determined rate law is rate = k[ICl][H ] . Write a two-step mechanism for this reaction using only
2

bimolecular elementary reactions and show that it is consistent with the experimental rate law. (Hint: HI is an intermediate.)

Answer
Solutions to Exercise 14.6.1
k1
step 1 rate = k1 [ICl][H2 ] (slow)
ICl + H2 −
→ HCl + HI

k2
step 2 HI + ICl −→ HCl + I2 rate = k2 [HI][ICl] (fast)
–––––––––––––––––––––

sum 2ICl + H2 → 2HCl + I2

This mechanism is consistent with the experimental rate law if the first step is the rate-determining step.

 Example 5.9.2 : Nitrogen Oxide Reacting with Molecular Hydrogen

Assume the reaction between NO and H occurs via a three-step process:


2

the reaction between NO and H occurs via a three-step process


2

k1
step 1 NO + NO −
→ N2 O2 (fast)

k2
step 2 N2 O2 + H2 −
→ N2 O + H2 O (slow)

k3
step 3 N2 O + H2 −
→ N2 + H2 O (fast)

Write the rate law for each elementary reaction, write the balanced chemical equation for the overall reaction, and identify the
rate-determining step. Is the rate law for the rate-determining step consistent with the experimentally derived rate law for the
overall reaction:
2
rate = k[NO] [ H ]? (observed)
2

5.9.4 https://chem.libretexts.org/@go/page/81885
Answer
Step 1: rate = k 1 [NO]
2

Step 2: rate = k 2 [ N2 O2 ][ H2 ]

Step 3: rate = k 3 [ N2 O][ H2 ]

The overall reaction is then

2 NO(g) + 2 H (g) ⟶ N (g) + 2 H O(g)


2 2 2

Rate Determining Step : #2


Yes, because the rate of formation of [N O ] = k [NO] . Substituting k [NO] for [N
2 2 1
2
1
2
2
O ]
2
in the rate law for step 2 gives
the experimentally derived rate law for the overall chemical reaction, where k = k k . 1 2

I'm not a robot


reCAPTCHA
Privacy - Terms

About this page

Our systems have detected unusual traffic from your


computer network This page checks to see if it's really you
Reaction Mechanism (Slow step followed by fast step): Reaction Mechanism (Slow step Followed by Fast Step)(opens in new
window) [youtu.be] (opens in new window)

Summary
A balanced chemical reaction does not necessarily reveal either the individual elementary reactions by which a reaction occurs or
its rate law. A reaction mechanism is the microscopic path by which reactants are transformed into products. Each step is an
elementary reaction. Species that are formed in one step and consumed in another are intermediates. Each elementary reaction can
be described in terms of its molecularity, the number of molecules that collide in that step. The slowest step in a reaction
mechanism is the rate-determining step.

5.9: Reaction Mechanisms is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
14.6: Reaction Mechanisms by Anonymous is licensed CC BY-NC-SA 3.0.

5.9.5 https://chem.libretexts.org/@go/page/81885
5.10: Catalysis
 Learning Objectives
To understand how catalysts increase the reaction rate and the selectivity of chemical reactions.

Catalysts are substances that increase the reaction rate of a chemical reaction without being consumed in the process. A catalyst,
therefore, does not appear in the overall stoichiometry of the reaction it catalyzes, but it must appear in at least one of the
elementary reactions in the mechanism for the catalyzed reaction. The catalyzed pathway has a lower Ea, but the net change in
energy that results from the reaction (the difference between the energy of the reactants and the energy of the products) is not
affected by the presence of a catalyst (Figure 5.10.1). Nevertheless, because of its lower Ea, the reaction rate of a catalyzed reaction
is faster than the reaction rate of the uncatalyzed reaction at the same temperature. Because a catalyst decreases the height of the
energy barrier, its presence increases the reaction rates of both the forward and the reverse reactions by the same amount. In this
section, we will examine the three major classes of catalysts: heterogeneous catalysts, homogeneous catalysts, and enzymes.

Figure 5.10.1 : Lowering the Activation Energy of a Reaction by a Catalyst. This graph compares potential energy diagrams for a
single-step reaction in the presence and absence of a catalyst. The only effect of the catalyst is to lower the activation energy of the
reaction. The catalyst does not affect the energy of the reactants or products (and thus does not affect ΔE). (CC BY-NC-SA;
anonymous)
The green line represents the uncatalyzed reaction. The purple line represent the catalyzed reaction .

A catalyst affects Ea, not ΔE.

Heterogeneous Catalysis
In heterogeneous catalysis, the catalyst is in a different phase from the reactants. At least one of the reactants interacts with the
solid surface in a physical process called adsorption in such a way that a chemical bond in the reactant becomes weak and then
breaks. Poisons are substances that bind irreversibly to catalysts, preventing reactants from adsorbing and thus reducing or
destroying the catalyst’s efficiency.
An example of heterogeneous catalysis is the interaction of hydrogen gas with the surface of a metal, such as Ni, Pd, or Pt. As
shown in part (a) in Figure 5.10.2, the hydrogen–hydrogen bonds break and produce individual adsorbed hydrogen atoms on the
surface of the metal. Because the adsorbed atoms can move around on the surface, two hydrogen atoms can collide and form a
molecule of hydrogen gas that can then leave the surface in the reverse process, called desorption. Adsorbed H atoms on a metal
surface are substantially more reactive than a hydrogen molecule. Because the relatively strong H–H bond (dissociation energy =
432 kJ/mol) has already been broken, the energy barrier for most reactions of H2 is substantially lower on the catalyst surface.

5.10.1 https://chem.libretexts.org/@go/page/81886
Figure 5.10.2 : Hydrogenation of Ethylene on a Heterogeneous Catalyst. When a molecule of hydrogen adsorbs to the catalyst
surface, the H–H bond breaks, and new M–H bonds are formed. The individual H atoms are more reactive than gaseous H2. When
a molecule of ethylene interacts with the catalyst surface, it reacts with the H atoms in a stepwise process to eventually produce
ethane, which is released. (CC BY-NC-SA; anonymous)
Figure 5.10.2 shows a process called hydrogenation, in which hydrogen atoms are added to the double bond of an alkene, such as
ethylene, to give a product that contains C–C single bonds, in this case ethane. Hydrogenation is used in the food industry to
convert vegetable oils, which consist of long chains of alkenes, to more commercially valuable solid derivatives that contain alkyl
chains. Hydrogenation of some of the double bonds in polyunsaturated vegetable oils, for example, produces margarine, a product
with a melting point, texture, and other physical properties similar to those of butter.
Several important examples of industrial heterogeneous catalytic reactions are in Table 5.10.1. Although the mechanisms of these
reactions are considerably more complex than the simple hydrogenation reaction described here, they all involve adsorption of the
reactants onto a solid catalytic surface, chemical reaction of the adsorbed species (sometimes via a number of intermediate species),
and finally desorption of the products from the surface.
Table 5.10.1 : Some Commercially Important Reactions that Employ Heterogeneous Catalysts
Commercial Process Catalyst Initial Reaction Final Commercial Product

contact process V2O5 or Pt 2SO2 + O2 → 2SO3 H2SO4

Haber process Fe, K2O, Al2O3 N2 + 3H2 → 2NH3 NH3

Ostwald process Pt and Rh 4NH3 + 5O2 → 4NO + 6H2O HNO3

H2 for NH3, CH3OH, and other


water–gas shift reaction Fe, Cr2O3, or Cu CO + H2O → CO2 + H2
fuels

steam reforming Ni CH4 + H2O → CO + 3H2 H2

methanol synthesis ZnO and Cr2O3 CO + 2H2 → CH3OH CH3OH

Sohio process bismuth phosphomolybdate CH 2 =CHCH3 + N H3 +


3

2
CH2 =CHCN
O2 → CH2 =CHCN + 3 H2 O
ac rylonitrile

RCH=CHR′ + H2 → RCH2— partially hydrogenated oils for


catalytic hydrogenation Ni, Pd, or Pt
CH2R′ margarine, and so forth

Homogeneous Catalysis
In homogeneous catalysis, the catalyst is in the same phase as the reactant(s). The number of collisions between reactants and
catalyst is at a maximum because the catalyst is uniformly dispersed throughout the reaction mixture. Many homogeneous catalysts
in industry are transition metal compounds (Table 5.10.2), but recovering these expensive catalysts from solution has been a major
challenge. As an added barrier to their widespread commercial use, many homogeneous catalysts can be used only at relatively low
temperatures, and even then they tend to decompose slowly in solution. Despite these problems, a number of commercially viable
processes have been developed in recent years. High-density polyethylene and polypropylene are produced by homogeneous
catalysis.
Table 5.10.2 : Some Commercially Important Reactions that Employ Homogeneous Catalysts
Commercial Process Catalyst Reactants Final Product

Union Carbide [Rh(CO)2I2]− CO + CH3OH CH3CO2H

5.10.2 https://chem.libretexts.org/@go/page/81886
Commercial Process Catalyst Reactants Final Product

hydroperoxide process Mo(VI) complexes CH3CH=CH2 + R–O–O–H

hydroformylation Rh/PR3 complexes RCH=CH2 + CO + H2 RCH2CH2CHO

NCCH2CH2CH2CH2CN used to
adiponitrile process Ni/PR3complexes 2HCN + CH2=CHCH=CH2
synthesize nylon
–(CH2CH2–)n: high-density
olefin polymerization (RC5H5)2ZrCl2 CH2=CH2
polyethylene

Enzymes
Enzymes, catalysts that occur naturally in living organisms, are almost all protein molecules with typical molecular masses of
20,000–100,000 amu. Some are homogeneous catalysts that react in aqueous solution within a cellular compartment of an
organism. Others are heterogeneous catalysts embedded within the membranes that separate cells and cellular compartments from
their surroundings. The reactant in an enzyme-catalyzed reaction is called a substrate.
Because enzymes can increase reaction rates by enormous factors (up to 1017 times the uncatalyzed rate) and tend to be very
specific, typically producing only a single product in quantitative yield, they are the focus of active research. At the same time,
enzymes are usually expensive to obtain, they often cease functioning at temperatures greater than 37 °C, have limited stability in
solution, and have such high specificity that they are confined to turning one particular set of reactants into one particular product.
This means that separate processes using different enzymes must be developed for chemically similar reactions, which is time-
consuming and expensive. Thus far, enzymes have found only limited industrial applications, although they are used as ingredients
in laundry detergents, contact lens cleaners, and meat tenderizers. The enzymes in these applications tend to be proteases, which
are able to cleave the amide bonds that hold amino acids together in proteins. Meat tenderizers, for example, contain a protease
called papain, which is isolated from papaya juice. It cleaves some of the long, fibrous protein molecules that make inexpensive
cuts of beef tough, producing a piece of meat that is more tender. Some insects, like the bombadier beetle, carry an enzyme capable
of catalyzing the decomposition of hydrogen peroxide to water (Figure 5.10.3).

Figure 5.10.3 : A Catalytic Defense Mechanism. The scalding, foul-smelling spray emitted by this bombardier beetle is produced
by the catalytic decomposition of H O .
2 2

Enzyme inhibitors cause a decrease in the reaction rate of an enzyme-catalyzed reaction by binding to a specific portion of an
enzyme and thus slowing or preventing a reaction from occurring. Irreversible inhibitors are therefore the equivalent of poisons in
heterogeneous catalysis. One of the oldest and most widely used commercial enzyme inhibitors is aspirin, which selectively
inhibits one of the enzymes involved in the synthesis of molecules that trigger inflammation. The design and synthesis of related
molecules that are more effective, more selective, and less toxic than aspirin are important objectives of biomedical research.

Summary
Catalysts participate in a chemical reaction and increase its rate. They do not appear in the reaction’s net equation and are not
consumed during the reaction. Catalysts allow a reaction to proceed via a pathway that has a lower activation energy than the
uncatalyzed reaction. In heterogeneous catalysis, catalysts provide a surface to which reactants bind in a process of adsorption. In
homogeneous catalysis, catalysts are in the same phase as the reactants. Enzymes are biological catalysts that produce large
increases in reaction rates and tend to be specific for certain reactants and products. The reactant in an enzyme-catalyzed reaction is
called a substrate. Enzyme inhibitors cause a decrease in the reaction rate of an enzyme-catalyzed reaction.

5.10.3 https://chem.libretexts.org/@go/page/81886
5.10: Catalysis is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
14.7: Catalysis is licensed CC BY-NC-SA 3.0.

5.10.4 https://chem.libretexts.org/@go/page/81886
5.11: Exercises

A general chemistry Libretexts Textmap organized around the textbook

 Chemistry: The Central Science


by Brown, LeMay, Bursten, Murphy, and Woodward

These are homework exercises to accompany the Textmap created for "Chemistry: The Central Science" by Brown et al.
Complementary General Chemistry question banks can be found for other Textmaps and can be accessed here. In addition to these
publicly available questions, access to private problems bank for use in exams and homework is available to faculty only on an
individual basis; please contact Delmar Larsen for an account with access permission.

14.1: Factors that Affect Reaction Rates


Q14.1.1
What information can you obtain by studying the chemical kinetics of a reaction? Does a balanced chemical equation provide the
same information? Why or why not?

S14.1.1
Kinetics gives information on the reaction rate and reaction mechanism; the balanced chemical equation gives only the
stoichiometry of the reaction.

Q14.1.2
If you were tasked with determining whether to proceed with a particular reaction in an industrial facility, why would studying the
chemical kinetics of the reaction be important to you?

S14.1.2
Studying chemical kinetics determines whether to proceed with a reaction as it measures the rate of a reaction. Reactions conducted
in an industrial facility mix compounds together, heating and stirring them for a while, before moving to the next process. When
the compounds are then moved to the next phase of the process it is important to know how long to hold the reaction at one stage
before continuing, to make sure a reaction has finished before starting the next one

Q14.1.3
What is the relationship between each of the following factors and the reaction rate: reactant concentration, temperature of the
reaction, physical properties of the reactants, physical and chemical properties of the solvent, and the presence of a catalyst?

S14.1.3
Reactant concentration:
Reaction rate and the concentration of reactants have a direct relationship; as concentration goes up, so does reaction rate. This is
because a higher concentration causes the reactants to have a higher probability to collide with a another reactant particle, possibly
inducing the chemical reaction. This comes from the Collision theory.
Reaction temperature:
Reaction rate and the reaction temperature also have a direct relationship; as temperature goes up, so does reaction rate. A higher
temperature makes the particles move at a faster speed so when two or more reactant particles collide, they collide with more
energy and are more likely to reach the activation energy threshold, starting a chemical reaction.
Physical properties of the reactants:
If the reactants have the same physical properties, they are more likely to react, increasing the reaction rate, because the reactants
mix and are more likely to collide another one of the reactants.
Physical and chemical properties of the solvent:

5.11.1 https://chem.libretexts.org/@go/page/81887
The properties of the solvent also affect the rates of a reaction. High viscosity means that particles more slowly than in low
viscosity solvents, so reaction rates are slower in high viscosity solvents than in low viscosity solvents. This indicates an inverse
relationship between viscosity and reaction rate.
Presence of a catalyst:
The chemical definition of a catalyst in a substance that increases the rate of a reaction without being used up in the reaction.
Therefore, if a catalyst is present, reaction rate increases.

Q14.1.4
A slurry is a mixture of a finely divided solid with a liquid in which it is only sparingly soluble. As you prepare a reaction, you
notice that one of your reactants forms a slurry with the solvent, rather than a solution. What effect will this have on the reaction
rate? What steps can you take to try to solve the problem?

Q14.1.5
Why does the reaction rate of virtually all reactions increase with an increase in temperature? If you were to make a glass of
sweetened iced tea the old-fashioned way, by adding sugar and ice cubes to a glass of hot tea, which would you add first?

S14.1.5
Increasing the temperature increases the average kinetic energy of molecules and ions, causing them to collide more frequently and
with greater energy, which increases the reaction rate. First dissolve sugar in the hot tea, and then add the ice.

Q14.1.6
In a typical laboratory setting, a reaction is carried out in a ventilated hood with air circulation provided by outside air. A student
noticed that a reaction that gave a high yield of a product in the winter gave a low yield of that same product in the summer, even
though his technique did not change and the reagents and concentrations used were identical. What is a plausible explanation for
the different yields?

Q14.1.7
A very active area of chemical research involves the development of solubilized catalysts that are not made inactive during the
reaction process. Such catalysts are expected to increase reaction rates significantly relative to the same reaction run in the presence
of a heterogeneous catalyst. What is the reason for anticipating that the relative rate will increase?

Q14.1.8
Water has a dielectric constant more than two times greater than that of methanol (80.1 for H2O and 33.0 for CH3OH). Which
would be your solvent of choice for a substitution reaction between an ionic compound and a polar reagent, both of which are
soluble in either methanol or water? Why?

14.2: Reaction Rates


Q14.2.1
Explain why the reaction rate is generally fastest at early time intervals. For the second-order A + B → C, what would the plot of
the concentration of C versus time look like during the course of the reaction?

S14.2.1
Reactant concentrations are highest at the beginning of a reaction. The plot of [C] versus t is a curve with a slope that becomes
steadily less positive.

Q14.2.2
Explain the differences between a differential rate law and an integrated rate law. What two components do they have in common?
Which form is preferred for obtaining a reaction order and a rate constant? Why?

Q14.2.3
Diffusion-controlled reactions have rates that are determined only by the reaction rate at which two reactant molecules can diffuse
together. These reactions are rapid, with second-order rate constants typically on the order of 1010 L/(mol·s). Are the reactions

5.11.2 https://chem.libretexts.org/@go/page/81887
expected to be faster or slower in solvents that have a low viscosity? Why? Consider the reactions H3O+ + OH− → 2H2O and H3O+
+ N(CH3)3 → H2O + HN(CH3)3+ in aqueous solution. Which would have the higher rate constant? Why?

S14.2.3
Faster in a less viscous solvent because the rate of diffusion is higher; the H3O+/OH− reaction is faster due to the decreased relative
size of reactants and the higher electrostatic attraction between the reactants.

Q14.2.4
What information can be obtained from the reaction order? What correlation does the reaction order have with the stoichiometry of
the overall equation?

Q14.2.5
During the hydrolysis reaction A + H2O → B + C, the concentration of A decreases much more rapidly in a polar solvent than in a
nonpolar solvent. How would this effect be reflected in the overall reaction order?

Q14.2.6
The reaction rate of a particular reaction in which A and B react to make C is as follows:
Δ[A] 1 Δ[C]
rate = − = ( )
Δt 2 Δt

Write a reaction equation that is consistent with this rate law. What is the rate expression with respect to time if 2A are converted to
3C?

Q14.2.7
While commuting to work, a person drove for 12 min at 35 mph, then stopped at an intersection for 2 min, continued the commute
at 50 mph for 28 min, drove slowly through traffic at 38 mph for 18 min, and then spent 1 min pulling into a parking space at 3
mph. What was the average rate of the commute? What was the instantaneous rate at 13 min? at 28 min?

Q14.2.8
Why do most studies of chemical reactions use the initial rates of reaction to generate a rate law? How is this initial rate
determined? Given the following data, what is the reaction order? Estimate.

Time (s) [A] (M)

120 0.158

240 0.089

360 0.062

Q14.2.9
Predict how the reaction rate will be affected by doubling the concentration of the first species in each equation.
a. C2H5I → C2H4 + HI: rate = k[C2H5I]
b. SO + O2 → SO2 + O: rate = k[SO][O2]
c. 2CH3 → C2H6: rate = k[CH3]2
d. ClOO → Cl + O2: rate = k

Q14.2.10
Cleavage of C2H6 to produce two CH3· radicals is a gas-phase reaction that occurs at 700°C. This reaction is first order, with k =
5.46 × 10−4 s−1. How long will it take for the reaction to go to 15% completion? to 50% completion?

S14.2.10
298 s; 1270 s

5.11.3 https://chem.libretexts.org/@go/page/81887
Q14.2.11
Three chemical processes occur at an altitude of approximately 100 km in Earth’s atmosphere.
k1
+ +
N + O2 −
→ N2 + O (5.11.1)
2 2

k2
+ +
O +O−
→ O2 + O (5.11.2)
2

k3
+ +
(O + N2 −
→ NO +N (5.11.3)

Write a rate law for each elementary reaction. If the rate law for the overall reaction were found to be rate = k[N2+][O2], which one
of the steps is rate limiting?

Q14.2.12
The oxidation of aqueous iodide by arsenic acid to give I3− and arsenous acid proceeds via the following reaction:
kf
− + −
H3 AsO4 (aq) + 3 I (aq) + 2 H (aq) ⇌ H3 AsO3 (aq) + I (aq) + H2 O(l)
3
kr

Write an expression for the initial rate of decrease of [I3−], Δ[I3−]/Δt. When the reaction rate of the forward reaction is equal to that
of the reverse reaction: kf/kr = [H3AsO3][I3−]/[H3AsO4][I−]3[H+]2. Based on this information, what can you say about the nature of
the rate-determining steps for the reverse and the forward reactions?

14.3: Concentration and Rate

14.4: The Change of Concentration with Time


Q14.4.1
What are the characteristics of a zeroth-order reaction? Experimentally, how would you determine whether a reaction is zeroth
order?

Q14.4.2
Predict whether the following reactions are zeroth order and explain your reasoning.
a. a substitution reaction of an alcohol with HCl to form an alkyl halide and water
b. catalytic hydrogenation of an alkene
c. hydrolysis of an alkyl halide to an alcohol
d. enzymatic conversion of nitrate to nitrite in a soil bacterium

Q14.4.3
In a first-order reaction, what is the advantage of using the integrated rate law expressed in natural logarithms over the rate law
expressed in exponential form?

Q14.4.4
If the reaction rate is directly proportional to the concentration of a reactant, what does this tell you about (a) the reaction order
with respect to the reactant and (b) the overall reaction order?

Q14.4.5
The reaction of NO with O2 is found to be second order with respect to NO and first order with respect to O2. What is the overall
reaction order? What is the effect of doubling the concentration of each reagent on the reaction rate?

Q14.4.6
Iodide reduces Fe(III) according to the following reaction:
3+ − 2+
2F e (soln) + 2 I (soln) → 2F e (soln) + I2 (soln) (5.11.4)

Experimentally, it was found that doubling the concentration of Fe(III) doubled the reaction rate, and doubling the iodide
concentration increased the reaction rate by a factor of 4. What is the reaction order with respect to each species? What is the

5.11.4 https://chem.libretexts.org/@go/page/81887
overall rate law? What is the overall reaction order?

S14.4.6
First order in Fe3+; second order in I−; third order overall; rate = k[Fe3+][I−]2.

Q14.4.7
Benzoyl peroxide is a medication used to treat acne. Its rate of thermal decomposition at several concentrations was determined
experimentally, and the data were tabulated as follows:

Experiment [Benzoyl Peroxide]0 (M) Initial Rate (M/s)

1 1.00 2.22 × 10−4

2 0.70 1.64 × 10−4

3 0.50 1.12 × 10−4

4 0.25 0.59 × 10−4

What is the reaction order with respect to benzoyl peroxide? What is the rate law for this reaction?

S14.4.7
The general rate law is: rate = k [Benzoyl Peroxide]m. In order to find the reaction order with respect to benzoyl peroxide, divide
two rate laws and solve for m:
m
rate2 k2 [BenzoylP eroxide]2
= ( )
rate1 k1 [BenzoylP eroxide]1

M

1.64 × 10 4
s 0.070M
m
=( )
M 1.00M

2.22 × 10 4
s

m
0.738 = (0.7)

m = 0.85

0.85
rate law: rate = k [Benzoyl Peroxide]

Q14.4.8
1-Bromopropane is a colorless liquid that reacts with S2O32− according to the following reaction:
2− − −
C3 H7 Br + S2 O → C3 H7 S2 O + Br (5.11.5)
3 3

The reaction is first order in 1-bromopropane and first order in S2O32−, with a rate constant of 8.05 × 10−4 M−1·s−1. If you began a
reaction with 40 mmol/100 mL of C3H7Br and an equivalent concentration of S2O32−, what would the initial reaction rate be? If
you were to decrease the concentration of each reactant to 20 mmol/100 mL, what would the initial reaction rate be?

S14.4.8
1.29 × 10−4 M/s; 3.22 × 10−5 M/s

Q14.4.9
The experimental rate law for the reaction 3A + 2B → C + D was found to be Δ[C]/Δt = k[A]2[B] for an overall reaction that is
third order. Because graphical analysis is difficult beyond second-order reactions, explain the procedure for determining the rate
law experimentally.

14.5: Temperature and Rate


Q14.5.1
Although an increase in temperature results in an increase in kinetic energy, this increase in kinetic energy is not sufficient to
explain the relationship between temperature and reaction rates. How does the activation energy relate to the chemical kinetics of a

5.11.5 https://chem.libretexts.org/@go/page/81887
reaction? Why does an increase in temperature increase the reaction rate despite the fact that the average kinetic energy is still less
than the activation energy?

S14.5.1
Activation energy is required for a collision between molecules to result in a chemical reaction, as well it is related to the rate due
to the Arrhenius equation
Activation energy is the threshold of energy needed in order for a reaction to occur. Reactant particles must collide with enough
energy to be able to break chemical bonds, that will then allow the creation of new bonds. If the particles do not have enough
energy, when they collide they will simply bounce off one another.
k=Ae(-Ea/RT). The increase in temperature increases the rate of reaction despite the fact the average kinetic is less than the activation
energy since it increase the rate of collision between molecules.
With an increase in temperature, there is a greater distribution of kinetic energy among reactant particles. This increase of
temperature allows the rate in which particles collide with one another to increase. Although the average kinetic energy is still
lower than the activation energy, the increase of collisions among particles increases the chance of particles that contain enough
energy to overcome the energy barrier to collide. Thus, the reaction rate increases due to this increased rate of collisions.
Additionally, there is a higher amount of molecules that have sufficient kinetic energy to overcome the energy barrier, despite the
average energy of all these molecules still being lower than the activation energy.

Q14.5.2
For any given reaction, what is the relationship between the activation energy and each of the following?
a. electrostatic repulsions
b. bond formation in the activated complex
c. the nature of the activated complex

S14.5.2
1.) Electrostatic repulsion: Electrostatic repulsion is the unfavorable interaction between two species of like charge. Activation
energy is the minimum amount of energy needed for a reaction to occur. Reacting molecules must have enough energy to overcome
electrostatic repulsion, and a minimum amount of energy is required to break chemical bonds so that new ones may be formed.
2.) Bond formation in the activated complex: An activated complex is an intermediate state that is formed during the conversion of
reactants into products. Bond breaking can increase activation energy as breaking bonds requires energy.
3.) Nature of the activated complex: The activation energy of a chemical reaction is the difference between the energy of the
activated complex and the energy of the reactants. If the structure has a high steric hindrance the activation energy will be higher.

Q14.5.3
If you are concerned with whether a reaction will occur rapidly, why would you be more interested in knowing the magnitude of
the activation energy than the change in potential energy for the reaction?

Q14.5.4
The product C in the reaction A + B → C + D can be separated easily from the reaction mixture. You have been given pure A and
pure B and are told to determine the activation energy for this reaction to determine whether the reaction is suitable for the
industrial synthesis of C. How would you do this? Why do you need to know the magnitude of the activation energy to make a
decision about feasibility?

Q14.5.5
Above Ea, molecules collide with enough energy to overcome the energy barrier for a reaction. Is it possible for a reaction to occur
at a temperature less than that needed to reach Ea? Explain your answer.

Q14.5.6
What is the relationship between A, Ea, and T? How does an increase in A affect the reaction rate?

5.11.6 https://chem.libretexts.org/@go/page/81887
Q14.5.7
Of two highly exothermic reactions with different values of Ea, which would need to be monitored more carefully: the one with the
smaller value or the one with the higher value? Why?

Q14.5.8
What happens to the approximate rate of a reaction when the temperature of the reaction is increased from 20°C to 30°C? What
happens to the reaction rate when the temperature is raised to 70°C? For a given reaction at room temperature (20°C), what is the
shape of a plot of reaction rate versus temperature as the temperature is increased to 70°C?

S14.5.8
The reaction rate will approximately double: 20°C to 30°C, the reaction rate increases by about 21 = 2; 20°C to 70°C, the reaction
rate increases by about 25 = 32-fold. A plot of reaction rate versus temperature will give an exponential increase: rate ∝ 2ΔT/10.

Q14.5.9
Acetaldehyde, used in silvering mirrors and some perfumes, undergoes a second-order decomposition between 700 and 840 K.
From the data in the following table, would you say that acetaldehyde follows the general rule that each 10 K increase in
temperature doubles the reaction rate?

T (K) k (M−1·s−1)

720 0.024

740 0.051

760 0.105

800 0.519

Q14.5.10
Bromoethane reacts with hydroxide ion in water to produce ethanol. The activation energy for this reaction is 90 kJ/mol. If the
reaction rate is 3.6 × 10−5 M/s at 25°C, what would the reaction rate be at the following temperatures?
a. 15°C
b. 30°C
c. 45°C

Q14.5.11
An enzyme-catalyzed reaction has an activation energy of 15 kcal/mol. How would the value of the rate constant differ between
20°C and 30°C? If the enzyme reduced the E from 25 kcal/mol to 15 kcal/mol, by what factor has the enzyme increased the
a

reaction rate at each temperature?

S14.5.11
a. 1.0 × 10−5 M/s
b. 6.6 × 10−5 M/s
c. 3.5 × 10−4 M/s

Q14.5.12
The data in the following table are the rate constants as a function of temperature for the dimerization of 1,3-butadiene. What is the
activation energy for this reaction?

T (K) k (M−1·min−1)

529 1.4

560 3.7

600 25

645 82

5.11.7 https://chem.libretexts.org/@go/page/81887
S14.5.12
100 kJ/mol

Q14.5.13
The reaction rate at 25°C is 1.0 × 10−4 M/s. Increasing the temperature to 75°C causes the reaction rate to increase to 7.0 × 10−2
M/s. Estimate E for this process. If E were 25 kJ/mol and the reaction rate at 25°C is 1.0 × 10−4 M/s, what would be the reaction
a a

rate at 75°C?

14.6: Reaction Mechanisms


Q14.6.1
How does the term molecularity relate to elementary reactions? How does it relate to the overall balanced chemical equation?

Q14.6.2
What is the relationship between the reaction order and the molecularity of a reaction? What is the relationship between the
reaction order and the balanced chemical equation?

Q14.6.3
When you determine the rate law for a given reaction, why is it valid to assume that the concentration of an intermediate does not
change with time during the course of the reaction?

Q14.6.3
If you know the rate law for an overall reaction, how would you determine which elementary reaction is rate determining? If an
intermediate is contained in the rate-determining step, how can the experimentally determined rate law for the reaction be derived
from this step?

Q14.6.4
Give the rate-determining step for each case.
a. Traffic is backed up on a highway because two lanes merge into one.
b. Gas flows from a pressurized cylinder fitted with a gas regulator and then is bubbled through a solution.
c. A document containing text and graphics is downloaded from the Internet.

Q14.6.5
Before being sent on an assignment, an aging James Bond was sent off to a health farm where part of the program’s focus was to
purge his body of radicals. Why was this goal considered important to his health?

S14.6.5
Free radicals are uncharged molecules with an unpaired valence electron. The reason these are so dangerous is because they like to
grab electrons from other atoms to fill their own outer shell. This allows them to impair protein function because free radicals
readily oxidize proteins and cell membrane which could lead to a loss of function. It was important to purge James Bond of radicals
because radicals set off chain reactions of continuously pulling electrons from molecules, which in turn, can damage cells in the
body.

Q14.6.6
Cyclopropane, a mild anesthetic, rearranges to propylene via a collision that produces and destroys an energized species. The
important steps in this rearrangement are as follows:

5.11.8 https://chem.libretexts.org/@go/page/81887
where M is any molecule, including cyclopropane. Only those cyclopropane molecules with sufficient energy (denoted with an
asterisk) can rearrange to propylene. Which step determines the rate constant of the overall reaction?

S14.6.6
When we are trying to decide which step is a rate-determining step, you have to look at which step is the slowest of them all, since
we have no knowledge of the time scale for either of these steps we have to look elsewhere to determine the rate-determining step.
In our specific problem, we do not know what exact molecules we have however we do see that the first step is an equilibrium
since it has a double-headed arrow with two k values one that is the reciprocal of the other. Since it is an equilibrium we can
reasonably assume that energized propane is rapidly created which would lead us to the conclusion that the second step is the
slower of the two steps and is, therefore, the rate-determining step. So, the second step (with k2) determines the rate constant of the
overall reaction.

Q14.6.7
Above approximately 500 K, the reaction between NO2 and CO to produce CO2 and NO follows the second-order rate law Δ[CO2]/
Δt = k[NO2][CO]. At lower temperatures, however, the rate law is Δ[CO2]/Δt = k′[NO2]2, for which it is known that NO3 is an
intermediate in the mechanism. Propose a complete low-temperature mechanism for the reaction based on this rate law. Which step
is the slowest?

S14.6.7
Given that NO3 is an intermediate in the low temperature reaction mechanism, we automatically know two things: 1) NO3 won't
show up in the final overall reaction, and thus, 2) NO3 will be in the products of the first reaction of the mechanism and in the
reactants of the second reaction. We can also assume that since we're given an intermediate from the problem, this is the only
intermediate (so we won't have to dream up any other compounds that might exist in the series of reactions.) These things being
said, the lower temperature reaction mechanism will look like this:
2N O2 (g) → N O3 (g) + N O(g) (1)

C O(g) + N O3 (g) → C O2 (g) + N O2 (g) (2)

And the overall reaction will look like this (notice how NO3 is not present):
C O(g) + N O2 (g) → C O2 (g) + N O(g) (overall reaction)

Now that we have the reaction mechanism written out, we can go about determining which step is the slowest. It would be pretty
tricky to do this if we weren't given any further information, however, we know two more things: 1) the high-temperature
mechanism's rate = k[NO2][CO] meaning that it took place in one step (given that the overall reaction is also equal to this rate)
and 2) the low-temperature mechanism's rate = k′[NO2]2. These two things being said, we've both confirmed that our proposed
low-temperature reaction mechanism is in fact two steps, and that we have a means to find which step is slower.
By using the "guess-and-check" method we can label each step reaction one at a time as the "slow reaction" and see if the rate
matches up with the rate given to us.
Let's first try the 2nd reaction. (see above)
By using rate laws we can determine that the rate of the reaction must be in terms of its reactants, which follows:
k[CO][NO ] (5.11.6)
3

... But wait! We can't have the overall reaction rate in terms of an intermediate.
By looking at the 1st reaction, we can determine that we can sub in "[NO2] /[NO]" for "[NO3]" since by writing the full reaction
rate of the first step and solving for [NO3] this is equivalent. So, we now have the overall rate of the mechanism as the following
given the 2nd reaction is the "slow reaction:"
2
[CO][NO ]
2
k (5.11.7)
[NO]

Note that NO2 is raised to the second power to account for the stoichiometry of the balanced reaction.
So clearly the 2nd reaction isn't the slow reaction since the rate is not equivalent to what we were given!

5.11.9 https://chem.libretexts.org/@go/page/81887
Let's check the rate of the 1st reaction now...
2
k[NO ] (5.11.8)
2

What do you know... the rates are equal!


We have now confirmed that the 1st reaction is the slow reaction equation, since its rate is equivalent to the overall reaction
rate.

Q14.6.8
Nitramide (O2NNH2) decomposes in aqueous solution to N2O and H2O. What is the experimental rate law (Δ[N2O]/Δt) for the
decomposition of nitramide if the mechanism for the decomposition is as follows?
k1
− +
O2 NN H2 ⇌ O2 NN H +H (fast)
k− 1

k2
− − (slow)
O2 NN H −
→ N2 O + OH

k3
+ − (fast)
H + OH −
→ H2 O

Assume that the rates of the forward and reverse reactions in the first equation are equal.

S14.6.8

We know that the slowest step of the reaction is the rate determining step, since it usually has the highest
activation energy requirement. As a result, the slowest step of the reaction is the experimental rate law
we are looking for.
Note: since the slowest step is the rate determining step, that usually means there is some intermediate in
between. Intermediates should NEVER be a part of the rate law mechanism.

rate = rate2 = k2 [ O2 N N H ] (5.11.9)

Since Nitramide is an intermediate, we must find some way to substitute it. To solve that problem, we
look for where Nitramide is produced and consumed. We see that
rate1 = k1 [ O2 N N H2 ] (5.11.10)

− +
rate−1 = k−1 [ O2 N N H ][ H ] (5.11.11)

Since these two rates produce and consume the same amount of O 2N NH

over the same time period, we can set them equal to
each other and solve for the intermediate

rate1 = rate−1 (5.11.12)

− +
k1 [ O2 N N H2 ] = k−1 [ O2 N N H ][ H ] (5.11.13)

k1 [ O2 N N H2 ]

[ O2 N N H ] = (5.11.14)
+
k−1 [ H ]

Substituting this equation back into our original equation gives us


k1 [ O2 N N H2 ]
rate = rate2 = k2 (5.11.15)
+
k−1 [ H ]

With all of the rate constants (k), we can clean up our equation a little bit by saying
k2 k1
k = (5.11.16)
k−1

Leaving us with

5.11.10 https://chem.libretexts.org/@go/page/81887
k[ O2 N N H2 ]
rate = (5.11.17)
+
[H ]

A14.6.8
k1 [ O2 NNH2 ] [ O2 NNH2 ]
rate = k2 =k
+ +
k−1 [ H ] [H ]

Q14.6.9
The following reactions are given:
k1

A+B ⇌ C +D (5.11.18)
k−1

k2

D+E−
→F (5.11.19)

What is the relationship between the relative magnitudes of k −1 and k if these reactions have the following rate law?
2

Δ[F ] [A][B][E]
=k (5.11.20)
Δt [C ]

How does the magnitude of k compare to that of k ? Under what conditions would you expect the rate law to be
1 2

Δ[F ]
= k'[A][B]? (5.11.21)
Δt

Assume that the rates of the forward and reverse reactions in the first equation are equal.

S14.6.9
First, because we have broken the equations down into elementary steps we can write the rate laws for each step.
Step1:
k1

A+B−
→ C +D (5.11.22)

rate = k1 [A][B] (5.11.23)

Step 2:
k−1

C +D −
−→ A+B (5.11.24)

rate = k−1 [C ][D] (5.11.25)

Step 3:
k2

D+E −
→F (5.11.26)

rate = k2 [D][E] (5.11.27)

If we add a these steps together we see that we get overall reaction


A+B+E → C +F (5.11.28)

we can see that [D] is an intermediate and


k1 = k−1 (5.11.29)

Since we are not told which steps are fast or slow we need to use Steady State Approximation.
If the second step is the slower step (k-1>>k2) then our rate determining step would be
rate = k2 [D][E] (5.11.30)

5.11.11 https://chem.libretexts.org/@go/page/81887
Since we can only write rate laws in terms of products and reactants we have to rewrite this so that we are not including an
intermediate.
Assume: rate of [D] formation = rate of its disappearance

k1 [A][B] = k−1 [C ][D] + k2 [D][E] (5.11.31)

k1 [A][B] = [D](k−1 [C ] + k2 [E]) (5.11.32)

Solving for [D] we find that


k1 [A][B]
[D] = (5.11.33)
(k−1 [C ] + k2 [E])

now we can use this to substitute the intermediate [D] in the rate law to get an appropriate rate law.
k2 k1 [A][B][E]
rate = (5.11.34)
(k−1 [C ] + k2 [E])

because we had already established k-1>>k2 we can assume that


k−1 [C ] + k2 [E] ≈ k−1 [C ] (5.11.35)

this would give us the observed rate law


Δ[F ] k2 k1 [A][B][E]
= (5.11.36)
Δt k−1 [C ]

to make this clearer we can set


(k2 )(k1 )
k = (5.11.37)
(k−1 )

and we can then simplify it down to


Δ[F ] [A][B][E]
=k (5.11.38)
Δt [C ]

we can see that all of these rate constants are related by this ratio k2k1/k-1. Since k2 is our rate determining step k-1>>k2 and since
k1=k-1 then we can see that k1>>k2.
We would expect the rate law to be
Δ[F ]
= k'[A][B]? (5.11.39)
Δt

if the rate determining step aka the slowest step is that corresponding to (step 1) since the rate law for this step is k1[A][B] and this
is the exact the same as the rate law that we were given.

14.7: Catalysis
Q14.7.1
What effect does a catalyst have on the activation energy of a reaction? What effect does it have on the frequency factor (A)? What
effect does it have on the change in potential energy for the reaction?

S14.7.1
A catalyst lowers the activation energy of a reaction. Some catalysts can also orient the reactants and thereby increase the
frequency factor. Catalysts have no effect on the change in potential energy for a reaction.

Q14.7.2
How is it possible to affect the product distribution of a reaction by using a catalyst?

5.11.12 https://chem.libretexts.org/@go/page/81887
Q14.7.3
A heterogeneous catalyst works by interacting with a reactant in a process called adsorption. What occurs during this process?
Explain how this can lower the activation energy.

S14.7.3
In adsorption, a reactant binds tightly to a surface. Because intermolecular interactions between the surface and the reactant weaken
or break bonds in the reactant, its reactivity is increased, and the activation energy for a reaction is often decreased.

Q14.7.4
What effect does increasing the surface area of a heterogeneous catalyst have on a reaction? Does increasing the surface area affect
the activation energy? Explain your answer.

Q14.7.5
Identify the differences between a heterogeneous catalyst and a homogeneous catalyst in terms of the following.
a. ease of recovery
b. collision frequency
c. temperature sensitivity
d. cost

S14.7.4
a. Heterogeneous catalysts are easier to recover.
b. Collision frequency is greater for homogeneous catalysts.
c. Homogeneous catalysts are often more sensitive to temperature.
d. Homogeneous catalysts are often more expensive.

Q14.7.6
An area of intensive chemical research involves the development of homogeneous catalysts, even though homogeneous catalysts
generally have a number of operational difficulties. Propose one or two reasons why a homogenous catalyst may be preferred.

S14.7.6
Catalysts are compounds that, when added to chemical reactions, reduce the activation energy and increase the reaction rate. The
amount of a catalyst does not change during a reaction, as it is not consumed as part of the reaction process. Catalysts lower the
energy required to reach the transition state of the reaction, allowing more molecular interactions to achieve that state. However,
catalysts do not affect the degree to which a reaction progresses. In other words, though catalysts affect reaction kinetics, the
equilibrium state remains unaffected.
Catalysts can be classified into two types: homogenous and heterogeneous. Homogenous catalysts are those which exist in the
same phase (gas or liquid) as the reactants, while heterogeneous catalysts are not in the same phase as the reactants. Typically,
heterogeneous catalysis involves the use of solid catalysts placed in a liquid reaction mixture. For this question, we will be
discussing homogenous catalysts.
Most of the times, homogeneous catalysis involves the introduction of an aqueous phase catalyst into an aqueous solution of
reactants. One reason why homogeneous catalysts are preferred over heterogeneous catalysts because homogeneous catalysts mix
well in chemical reactions in comparison to heterogeneous catalysts. However, homogeneous catalyst is often irrecoverable after
the reaction has run to completion.

Q14.7.7
Consider the following reaction between cerium(IV) and thallium(I) ions:
4 + + 3 + 3 +
2 Ce + Tl → 2 Ce + Tl (5.11.40)

This reaction is slow, but Mn2+ catalyzes it, as shown in the following mechanism:
4 + 2 + 3 + 3 +
Ce + Mn → Ce + Mn (5.11.41)

4 + 3 + 3 + 4 +
Ce + Mn → Ce + Mn (5.11.42)

5.11.13 https://chem.libretexts.org/@go/page/81887
4 + + 3 + 2 +
Mn + Tl → Tl + Mn (5.11.43)

In what way does M n 2+


increase the reaction rate?

S14.7.7
The Mn2+ ion donates two electrons to Ce4+, one at a time, and then accepts two electrons from Tl+. Because Mn can exist in three
oxidation states separated by one electron, it is able to couple one-electron and two-electron transfer reactions.

Q14.7.8
The text identifies several factors that limit the industrial applications of enzymes. Still, there is keen interest in understanding how
enzymes work for designing catalysts for industrial applications. Why?

S14.7.8
Enzymes are expensive to make, denature and fail at certain temperatures, are not that stable in a solution, and are very specific to
the reaction it was made for. However, scientists can use the observations from enzymes to create catalysts that are more effective
in aiding the reaction and cost less to produce. Overall, catalysts still play a large part in lowering the activation energy for
reactions. Creating new catalysts can help in the improvement of areas such as medical, ecological, and even commercial products.

Q14.7.9
Most enzymes have an optimal pH range; however, care must be taken when determining pH effects on enzyme activity. A
decrease in activity could be due to the effects of changes in pH on groups at the catalytic center or to the effects on groups located
elsewhere in the enzyme. Both examples are observed in chymotrypsin, a digestive enzyme that is a protease that hydrolyzes
polypeptide chains. Explain how a change in pH could affect the catalytic activity due to (a) effects at the catalytic center and (b)
effects elsewhere in the enzyme. (Hint: remember that enzymes are composed of functional amino acids.)

Q14.7.10
At some point during an enzymatic reaction, the concentration of the activated complex, called an enzyme–substrate complex (
ES ), and other intermediates involved in the reaction is nearly constant. When a single substrate is involved, the reaction can be

represented by the following sequence of equations:


enzyme (E) + substrate (S) ⇌ enzyme-substrate complex (ES) ⇌ enzyme (E) + product (P) (5.11.44)

This can also be shown as follows:


k1 k2

E + S ⇌ ES ⇌ E + P (5.11.45)
k−1 k−2

Using molar concentrations and rate constants, write an expression for the rate of disappearance of the enzyme–substrate complex.
Typically, enzyme concentrations are small, and substrate concentrations are high. If you were determining the rate law by varying
the substrate concentrations under these conditions, what would be your apparent reaction order?

S14.7.10
Δ[ES]
= −(k2 + k−1 )[ES] + k1 [E][S] + k−2 [E][P ] ≈ 0 ; zeroth order in substrate.
Δt

Q14.7.11
A particular reaction was found to proceed via the following mechanism:
A+B → C +D (slow)
2C → E (fast)
E +A → B+F (fast)
What is the overall reaction? Is this reaction catalytic, and if so, what species is the catalyst? Identify the intermediates

5.11.14 https://chem.libretexts.org/@go/page/81887
Q14.7.12
A particular reaction has two accessible pathways (A and B), each of which favors conversion of X to a different product (Y and Z,
respectively). Under uncatalyzed conditions pathway A is favored, but in the presence of a catalyst pathway B is favored. Pathway
B is reversible, whereas pathway A is not. Which product is favored in the presence of a catalyst? without a catalyst? Draw a
diagram illustrating what is occurring with and without the catalyst.

S14.7.12
In both cases, the product of pathway A is favored. All of the Z produced in the catalyzed reversible pathway B will eventually be
converted to X as X is converted irreversibly to Y by pathway A.

Q14.7.13
The kinetics of an enzyme-catalyzed reaction can be analyzed by plotting the reaction rate versus the substrate concentration. This
type of analysis is referred to as a Michaelis–Menten treatment. At low substrate concentrations, the plot shows behavior
characteristic of first-order kinetics, but at very high substrate concentrations, the behavior shows zeroth-order kinetics. Explain
this phenomenon.

5.11: Exercises is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

5.11.15 https://chem.libretexts.org/@go/page/81887
5.12: Chemical Kinetics (Summary)

A general chemistry Libretexts Textmap organized around the textbook

 Chemistry: The Central Science


by Brown, LeMay, Bursten, Murphy, and Woodward

These is a summary of key concepts of the chapter in the Textmap created for "Chemistry: The Central Science" by Brown et al.

14.1: Factors that Affect Reaction Rates


chemical kinetics – area of chemistry dealing with speeds/rates of reactions
rates of reactions affected by four factors
1. concentrations of reactants
2. temperature at which reaction occurs
3. presence of a catalyst
4. surface area of solid or liquid reactants and/or catalysts

14.2: Reaction Rates


reaction rate – speed of a chemical reaction
change #moles B Δmoles B
average rate = =  if A → B
change in time Δt

Δmoles B = moles B at final time − moles B at initial time

Δmoles A
average rate = −  if A → B
Δt

14.2.1 Rates in Terms of Concentrations


rate calculated in units of M/s
brackets around a substance indicate the concentration
instantaneous rate – rate at a particular time
instantaneous rate obtained from the straight line tangent that touches the curve at a specific point
slopes give instantaneous rates
instantaneous rate also referred to as the rate

14.2.2 Reaction Rates and Stoichiometry


for the irreversible reaction aA + bB → cC + dD
1 Δ[A] 1 Δ[B] 1 Δ[C ] 1 Δ[D]
rate = − =− = =
a Δt b Δt c Δt d Δt

14.3: Concentration and Rate


equation used only if C and D only substances formed
Rate = k[A][B]
Rate law – expression that shows that rate depends on concentrations of reactants
k = rate constant

14.3.1 Reaction Order


Rate = k[reactant 1]m[reactant 2]n
m, n are called reaction orders
m+n, overall reaction order
reaction orders do not have to correspond with coefficients in balanced equation
values of reaction order determined experimentally

5.12.1 https://chem.libretexts.org/@go/page/81888
reaction order can be fractional or negative

14.3.2 Units of Rates Constants


units of rate constant depend on overall reaction order of rate law
for reaction of second order overall
units of rate = (units of rate constant)(units of concentration)2
units of rate constant = M-1s-1

14.3.3 Using Initial Rates to Determine Rate Laws


zero order – no change in rate when concentration changed
first order – change in concentration gives proportional changes in rate
second order – change in concentration changes rate by the square of the concentration change, such as 22 or 32, etc…
rate constant does not depend on concentration

14.4: The Change of Concentration with Time


rate laws can be converted into equations that give concentrations of reactants or products

14.4.1 First-Order Reactions


Δ[A]
rate = − = k[A]
Δt

and in integral form:

ln[A]t − ln[A]0 = −kt

or
[A]t
ln = −kt
[A]0

ln[A]t = −kt + ln[A]0

corresponds to a straight line with y = mx + b


equations used to determine:
1. concentration of reactant remaining at any time
2. time required for given fraction of sample to react
3. time required for reactant concentration to reach a certain level

14.3.2 Half-Life
half-life of first order reaction
1
ln 0.693
2
t 1 =− =
2 k k

half-life – time required for concentration of reactant to drop to one-half of initial value
t
1/2 of first order independent of initial concentrations
half-life same at any given time of reaction
in first order reaction – concentrations of reactant decreases by ½ in each series of regularly spaced time intervals

14.3.3 Second-Order Reactions


rate depends on reactant concentration raised to second power or concentrations of two different reactants each raised to first
power
2
Rate = k[A]

1 1
= kt +
[A]t [A]0

5.12.2 https://chem.libretexts.org/@go/page/81888
1
half life = t 1 =
2 k[A]0

half life dependent on initial concentration of reactant

14.5: Temperature and Rate


rate constant must increase with increasing temperature, thus increasing the rate of reaction

14.5.1 The Collision Model


collision model – molecules must collide to react
greater frequency of collisions the greater the reaction rate
for most reactions only a small fraction of collisions leads to a reaction

14.5.2 Activation Energy


Svante August Arrhenius
Molecules must have a minimum amount of energy to react
Energy comes from kinetic energy of collisions
Kinetic energy used to break bonds
Activation energy, Ea – minimum energy required to initiate a chemical reaction
Activated complex or transition state – atoms at the top of the energy barrier
Rate depends on temperature and Ea
Lower Ea means faster reaction
Reactions occur when collisions between molecules occur with enough energy and proper orientation

14.5.3 The Arrhenius Equation


reaction rate data:
theArrhenius Equation:
−Ea

k = Ae RT

k = rate constant, E = activation energy, R = gas constant (8.314 J/(mol K)), T = absolute temperature, A = frequency factor
a

A relates to frequency of collisions, favorable orientations


Ea
ln k = − + ln A
RT

the ln k vs. 1/t graph (also known as an Arrhenius plot) has a slope – E a /R and the y-intercept ln A
for two temperatures:
k1 Ea 1 1
ln = ( − )
k2 R T2 T1

used to calculate rate constant, k and T


1 1

14.6: Reaction Mechanisms


reaction mechanism – process by which a reaction occurs

14.6.1 Elementary Steps


elementary steps – each step in a reaction
molecularity – if only one molecule involved in step
unimolecular – if only one molecule involved in step
bimolecular – elementary step involving collision of two reactant molecules
termolecular – elementary step involving simultaneous collision of three molecules
elementary steps in multi-step mechanism must always add to give chemical equation of overall process
intermediate – product formed in one step and consumed in a later step

5.12.3 https://chem.libretexts.org/@go/page/81888
14.6.2 Rate Laws of Elementary Steps
if reaction is known to be an elementary step then the rate law is known
rate of unimolecular step is first order (Rate = k[A])
rate of bimolecular steps is second order (Rate = k[A][B])
first order in [A] and [B]
if double [A] than number of collisions of A and B will double

14.6.3 Rate Laws of Multi-step Mechanisms


rate-determining step – slowest elementary step
determines rate law of overall reaction

14.6.4 Mechanisms with an Initial Slow Step vs. Mechanisms with an Initial Fast Step
intermediates are usually unstable, in low concentration, and difficult to isolate
when a fast step precedes a slow one, solve for concentration of intermediate by assuming that equilibrium is established in fast
step

14.7: Catalysis
catalyst – substance that changes speed of chemical reaction without undergoing a permanent chemical change

14.7.1 Homogeneous Catalysis


homogeneous catalyst – catalyst that is present in same phase as reacting molecule
catalysts alter Ea or A
generally catalysts lowers overall Ea for chemical reaction
catalysts provides a different mechanism for reaction

14.7.2 Heterogeneous Catalysis


exists in different phase from reactants
initial step in heterogeneous catalyst is adsorption
adsorption – binding of molecules to surface
adsorption occurs because ions/atoms at surface of solid extremely reactive

14.7.3 Enzymes
biological catalysts
large protein molecules with molecular weights 10,000 – 1 million amu
catalase – enzyme in blood and liver that decomposes hydrogen peroxide into water and oxygen
substrates – substances that undergo reaction at the active site
lock-and-key model – substrate molecules bind specifically to the active site
enzyme-substrate complex – combination of enzyme and substrate
binding between enzyme and substrate involves intermolecular forces (dipole-dipole, hydrogen bonding, and London dispersion
forces)
product from reaction leaves enzyme allowing for another substrate to enter enzyme
enzyme inhibitors – molecules that bind strongly to enzymes
turnover number – number of catalyzed reactions occurring at a particular active site
large turnover numbers = low activation energies

5.12: Chemical Kinetics (Summary) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
14.S: Chemical Kinetics (Summary) is licensed CC BY-NC-SA 3.0.

5.12.4 https://chem.libretexts.org/@go/page/81888
CHAPTER OVERVIEW
6: Nuclear Chemistry
UC Davis - Spring 2021

Chemistry 2C
Unit 1: Electrochemistry  ●  Unit 2: Coordination Chemistry  ●  Unit 3: Descriptive Chemistry  ●  Unit 4: Chemical
Kinetics  ●  Unit 5: Nuclear Chemistry
  Agenda  ●  Textbook  ●  ADAPT  ●  Worksheets

Nuclear reactions differ from other chemical processes in one critical way: in a nuclear reaction, the identities of the elements
change. In addition, nuclear reactions are often accompanied by the release of enormous amounts of energy, as much as a billion
times more than the energy released by chemical reactions. Moreover, the yields and rates of a nuclear reaction are generally
unaffected by changes in temperature, pressure, or the presence of a catalyst.
6.1: Components of the Nucleus
6.2: Nuclear Reactions
6.3: Nuclear Radiation
6.4: Rates of Radioactive Decay
6.5: Stability of the Atomic Nucleus
6.6: The Origin of the Elements
6.7: Transmutation of the Elements
6.8: Nuclear Fission
6.9: Nuclear Fusion
6.10: Nuclear Chemistry (Exercises)

6: Nuclear Chemistry is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

1
6.1: Components of the Nucleus
24.03: Components of the Nucleus

6.1: Components of the Nucleus is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

6.1.1 https://chem.libretexts.org/@go/page/13190
6.2: Nuclear Reactions
 Learning Objectives
To know the different kinds of radioactive decay.
To balance a nuclear reaction.

The two general kinds of nuclear reactions are nuclear decay reactions and nuclear transmutation reactions. In a nuclear decay
reaction, also called radioactive decay, an unstable nucleus emits radiation and is transformed into the nucleus of one or more other
elements. The resulting daughter nuclei have a lower mass and are lower in energy (more stable) than the parent nucleus that
decayed. In contrast, in a nuclear transmutation reaction, a nucleus reacts with a subatomic particle or another nucleus to form a
product nucleus that is more massive than the starting material. As we shall see, nuclear decay reactions occur spontaneously under
all conditions, but nuclear transmutation reactions occur only under very special conditions, such as the collision of a beam of
highly energetic particles with a target nucleus or in the interior of stars. We begin this section by considering the different classes
of radioactive nuclei, along with their characteristic nuclear decay reactions and the radiation they emit.

Nuclear decay reactions occur spontaneously under all conditions, whereas nuclear
transmutation reactions are induced.

Classes of Radioactive Nuclei


The three general classes of radioactive nuclei are characterized by a different decay process or set of processes:
1. Neutron-rich nuclei. The nuclei on the upper left side of the band of stable nuclei have a neutron-to-proton ratio that is too
high to give a stable nucleus. These nuclei decay by a process that converts a neutron to a proton, thereby decreasing the
neutron-to-proton ratio.
2. Neutron-poor nuclei. Nuclei on the lower right side of the band of stable nuclei have a neutron-to-proton ratio that is too low
to give a stable nucleus. These nuclei decay by processes that have the net effect of converting a proton to a neutron, thereby
increasing the neutron-to-proton ratio.
3. Heavy nuclei. With very few exceptions, heavy nuclei (those with A ≥ 200) are intrinsically unstable regardless of the neutron-
to-proton ratio, and all nuclei with Z > 83 are unstable. This is presumably due to the cumulative effects of electrostatic
repulsions between the large number of positively charged protons, which cannot be totally overcome by the strong nuclear
force, regardless of the number of neutrons present. Such nuclei tend to decay by emitting an α particle (a helium nucleus, He,4
2

which decreases the number of protons and neutrons in the original nucleus by 2. Because the neutron-to-proton ratio in an α
particle is 1, the net result of alpha emission is an increase in the neutron-to-proton ratio.

Nuclear decay reactions always produce daughter nuclei that have a more favorable
neutron-to- proton ratio and hence are more stable than the parent nucleus.

Nuclear Decay Reactions


Just as we use the number and type of atoms present to balance a chemical equation, we can use the number and type of nucleons
present to write a balanced nuclear equation for a nuclear decay reaction. This procedure also allows us to predict the identity of
either the parent or the daughter nucleus if the identity of only one is known. Regardless of the mode of decay, the total number of
nucleons is conserved in all nuclear reactions.
To describe nuclear decay reactions, chemists have extended the X notation for nuclides to include radioactive emissions. Table
A
Z

6.2.1 lists the name and symbol for each type of emitted radiation. The most notable addition is the positron, a particle that has the

same mass as an electron but a positive charge rather than a negative charge.
Table 6.2.1 : Nuclear Decay Emissions and Their Symbols
Identity Symbol Charge Mass (amu)

helium nucleus 4
2
α +2 4.001506

electron 0
−1
β or β −
−1 0.000549

photon 0
0
γ — —

6.2.1 https://chem.libretexts.org/@go/page/13191
Identity Symbol Charge Mass (amu)

neutron 1
0
n 0 1.008665

proton 1
1
p +1 1.007276

positron 0
+1
β or β +
+1 0.000549

Like the notation used to indicate isotopes, the upper left superscript in the symbol for a particle gives the mass number, which is
the total number of protons and neutrons. For a proton or a neutron, A = 1. Because neither an electron nor a positron contains
protons or neutrons, its mass number is 0. The numbers should not be taken literally, however, as meaning that these particles have
zero mass; ejection of a beta particle (an electron) simply has a negligible effect on the mass of a nucleus.
Similarly, the lower left subscript gives the charge of the particle. Because protons carry a positive charge, Z = +1 for a proton. In
contrast, a neutron contains no protons and is electrically neutral, so Z = 0. In the case of an electron, Z = −1, and for a positron, Z =
+1. Because γ rays are high-energy photons, both A and Z are 0. In some cases, two different symbols are used for particles that are
identical but produced in different ways. For example, the symbol e, which is usually simplified to e−, represents a free electron
0
−1

or an electron associated with an atom, whereas the symbol β, which is often simplified to β−, denotes an electron that originates
0
−1

from within the nucleus, which is a β particle. Similarly, He refers to the nucleus of a helium atom, and α denotes an identical
4
2
2+ 4
2

particle that has been ejected from a heavier nucleus.


There are six fundamentally different kinds of nuclear decay reactions, and each releases a different kind of particle or energy. The
essential features of each reaction are shown in Figure 6.2.1. The most common are alpha and beta decay and gamma emission, but
the others are essential to an understanding of nuclear decay reactions.

Figure 6.2.1 : Common Modes of Nuclear Decay

6.2.2 https://chem.libretexts.org/@go/page/13191
Alpha Decay
Many nuclei with mass numbers greater than 200 undergo alpha (α) decay, which results in the emission of a helium-4 nucleus as
an alpha (α) particle, α . The general reaction is as follows:
4
2

A A−4 ′ 4
X → X + α (6.2.1)
Z Z−2 2
parent daughter alpha

particle

The daughter nuclide contains two fewer protons and two fewer neutrons than the parent. Thus α-particle emission produces a
daughter nucleus with a mass number A − 4 and a nuclear charge Z − 2 compared to the parent nucleus. Radium-226, for example,
undergoes alpha decay to form radon-222:
226 222 4
Ra → Rn + α (6.2.2)
88 86 2

Because nucleons are conserved in this and all other nuclear reactions, the sum of the mass numbers of the products, 222 + 4 = 226,
equals the mass number of the parent. Similarly, the sum of the atomic numbers of the products, 86 + 2 = 88, equals the atomic
number of the parent. Thus the nuclear equation is balanced.

Just as the total number of atoms is conserved in a chemical reaction, the total number of nucleons is conserved in a nuclear
reaction.

Beta Decay
Nuclei that contain too many neutrons often undergo beta (β) decay, in which a neutron is converted to a proton and a high-energy
electron that is ejected from the nucleus as a β particle:
1 1 0
n → p + β (6.2.3)
0 1 −1
unstable proton beta particle

neutron in retained emitted by

nucleus by nucleus nucleus

The general reaction for beta decay is therefore


A A ′ 0
X → X + β (6.2.4)
Z Z+1 −1
parent daughter beta particle

Although beta decay does not change the mass number of the nucleus, it does result in an increase of +1 in the atomic number
because of the addition of a proton in the daughter nucleus. Thus beta decay decreases the neutron-to-proton ratio, moving the
nucleus toward the band of stable nuclei. For example, carbon-14 undergoes beta decay to form nitrogen-14:
14 14 0
C → N+ β (6.2.5)
6 7 −1

Once again, the number of nucleons is conserved, and the charges are balanced. The parent and the daughter nuclei have the same
mass number, 14, and the sum of the atomic numbers of the products is 6, which is the same as the atomic number of the carbon-14
parent.

Positron Emission
Because a positron has the same mass as an electron but opposite charge, positron emission is the opposite of beta decay. Thus
positron emission is characteristic of neutron-poor nuclei, which decay by transforming a proton to a neutron and emitting a high-
energy positron:
1 + 1 0 +
p → n+ β (6.2.6)
1 0 +1

The general reaction for positron emission is therefore


A A ′ 0 +
X → X + β (6.2.7)
Z Z−1 +1
parent daughter positron

Like beta decay, positron emission does not change the mass number of the nucleus. In this case, however, the atomic number of
the daughter nucleus is lower by 1 than that of the parent. Thus the neutron-to-proton ratio has increased, again moving the nucleus
closer to the band of stable nuclei. For example, carbon-11 undergoes positron emission to form boron-11:

6.2.3 https://chem.libretexts.org/@go/page/13191
11 11 0 +
C → B+ β (6.2.8)
6 5 +1

Nucleons are conserved, and the charges balance. The mass number, 11, does not change, and the sum of the atomic numbers of the
products is 6, the same as the atomic number of the parent carbon-11 nuclide.

Electron Capture
A neutron-poor nucleus can decay by either positron emission or electron capture (EC), in which an electron in an inner shell
reacts with a proton to produce a neutron:
1 0 1
p+ e → n (6.2.9)
1 −1 0

When a second electron moves from an outer shell to take the place of the lower-energy electron that was absorbed by the nucleus,
an x-ray is emitted. The overall reaction for electron capture is thus
A 0 A ′
X + e → X + x-ray (6.2.10)
Z −1 Z−1
parent electron daughter

Electron capture does not change the mass number of the nucleus because both the proton that is lost and the neutron that is formed
have a mass number of 1. As with positron emission, however, the atomic number of the daughter nucleus is lower by 1 than that of
the parent. Once again, the neutron-to-proton ratio has increased, moving the nucleus toward the band of stable nuclei. For
example, iron-55 decays by electron capture to form manganese-55, which is often written as follows:
EC
55 55
Fe → Mn + x-ray (6.2.11)
26 25

The atomic numbers of the parent and daughter nuclides differ in Equation 20.2.11, although the mass numbers are the same. To
write a balanced nuclear equation for this reaction, we must explicitly include the captured electron in the equation:
55 0 55
Fe + e → Mn + x-ray (6.2.12)
26 −1 25

Both positron emission and electron capture are usually observed for nuclides with low neutron-to-proton ratios, but the decay rates
for the two processes can be very different.

Gamma Emission
Many nuclear decay reactions produce daughter nuclei that are in a nuclear excited state, which is similar to an atom in which an
electron has been excited to a higher-energy orbital to give an electronic excited state. Just as an electron in an electronic excited
state emits energy in the form of a photon when it returns to the ground state, a nucleus in an excited state releases energy in the
form of a photon when it returns to the ground state. These high-energy photons are γ rays. Gamma (γ) emission can occur
virtually instantaneously, as it does in the alpha decay of uranium-238 to thorium-234, where the asterisk denotes an excited state:
relaxation
238 234 4 234 0
U → Th* + α −−−−−−→ Th + γ (6.2.13)
92 90 2 90 0
excited

nuclear

state

If we disregard the decay event that created the excited nucleus, then
234 234 0
Th* → Th + γ (6.2.14)
88 88 0

or more generally,
A A 0
X* → X+ γ (6.2.15)
Z Z 0

Gamma emission can also occur after a significant delay. For example, technetium-99m has a half-life of about 6 hours before
emitting a γ ray to form technetium-99 (the m is for metastable). Because γ rays are energy, their emission does not affect either the
mass number or the atomic number of the daughter nuclide. Gamma-ray emission is therefore the only kind of radiation that does
not necessarily involve the conversion of one element to another, although it is almost always observed in conjunction with some
other nuclear decay reaction.

6.2.4 https://chem.libretexts.org/@go/page/13191
Spontaneous Fission
Only very massive nuclei with high neutron-to-proton ratios can undergo spontaneous fission, in which the nucleus breaks into
two pieces that have different atomic numbers and atomic masses. This process is most important for the transactinide elements,
with Z ≥ 104. Spontaneous fission is invariably accompanied by the release of large amounts of energy, and it is usually
accompanied by the emission of several neutrons as well. An example is the spontaneous fission of Cf , which gives a
254
98

distribution of fission products; one possible set of products is shown in the following equation:
254 118 132 1
Cf → Pd + Te + 40 n (6.2.16)
98 46 52

Once again, the number of nucleons is conserved. Thus the sum of the mass numbers of the products (118 + 132 + 4 = 254) equals
the mass number of the reactant. Similarly, the sum of the atomic numbers of the products [46 + 52 + (4 × 0) = 98] is the same as
the atomic number of the parent nuclide.

 Example 6.2.1
Write a balanced nuclear equation to describe each reaction.
a. the beta decay of S
35
16

b. the decay of Hg by electron capture


201
80

c. the decay of P by positron emission


30
15

Given: radioactive nuclide and mode of decay


Asked for: balanced nuclear equation
Strategy:
A Identify the reactants and the products from the information given.
B Use the values of A and Z to identify any missing components needed to balance the equation.
Solution
a.
A We know the identities of the reactant and one of the products (a β particle). We can therefore begin by writing an
equation that shows the reactant and one of the products and indicates the unknown product as X: A
Z

35 A 0
S→ X+ β (6.2.17)
16 Z −1

B Because both protons and neutrons must be conserved in a nuclear reaction, the unknown product must have a mass
number of A = 35 − 0 = 35 and an atomic number of Z = 16 − (−1) = 17. The element with Z = 17 is chlorine, so the
balanced nuclear equation is as follows:
35 35 0
S→ Cl + β (6.2.18)
16 17 −1

b.
A We know the identities of both reactants: 201
80
Hg and an inner electron, 0
−1
. The reaction is as follows:
e

201 0 A
Hg + e → X
80 −1 Z

B Both protons and neutrons are conserved, so the mass number of the product must be A = 201 + 0 = 201, and the
atomic number of the product must be Z = 80 + (−1) = 79, which corresponds to the element gold. The balanced nuclear
equation is thus
201 0 201
Hg + e → Au
80 −1 79

c.
A As in part (a), we are given the identities of the reactant and one of the products—in this case, a positron. The
unbalanced nuclear equation is therefore
30 A 0
P → X+ β
15 Z +1

6.2.5 https://chem.libretexts.org/@go/page/13191
B The mass number of the second product is A = 30 − 0 = 30, and its atomic number is Z = 15 − 1 = 14, which
corresponds to silicon. The balanced nuclear equation for the reaction is as follows:
30 30 0
P → Si + β
15 14 +1

 Exercise 6.2.1

Write a balanced nuclear equation to describe each reaction.


a. C by positron emission
11
6

b. the beta decay of molybdenum-99


c. the emission of an α particle followed by gamma emission from 185
74
W

Answer
a. 11
6
C →
11
5
B+
0
+1
β

b. 99
42
Mo →
99m
43
Tc +
0
−1
β

c. 185
74
W →
181
72
Hf +
4
2
α+
0
0
γ

 Example 6.2.2

Predict the kind of nuclear change each unstable nuclide undergoes when it decays.
a. 45
22
Ti

b. 242
94
Pu

c. 12
5
B

d. 256
100
Fm

Given: nuclide
Asked for: type of nuclear decay
Strategy:
Based on the neutron-to-proton ratio and the value of Z, predict the type of nuclear decay reaction that will produce a more
stable nuclide.
Solution
a. This nuclide has a neutron-to-proton ratio of only 1.05, which is much less than the requirement for stability for an element
with an atomic number in this range. Nuclei that have low neutron-to-proton ratios decay by converting a proton to a
neutron. The two possibilities are positron emission, which converts a proton to a neutron and a positron, and electron
capture, which converts a proton and a core electron to a neutron. In this case, both are observed, with positron emission
occurring about 86% of the time and electron capture about 14% of the time.
b. Nuclei with Z > 83 are too heavy to be stable and usually undergo alpha decay, which decreases both the mass number and
the atomic number. Thus Pu is expected to decay by alpha emission.
242
94

c. This nuclide has a neutron-to-proton ratio of 1.4, which is very high for a light element. Nuclei with high neutron-to-proton
ratios decay by converting a neutron to a proton and an electron. The electron is emitted as a β particle, and the proton
remains in the nucleus, causing an increase in the atomic number with no change in the mass number. We therefore predict
that B will undergo beta decay.
12
5

d. This is a massive nuclide, with an atomic number of 100 and a mass number much greater than 200. Nuclides with A ≥ 200
tend to decay by alpha emission, and even heavier nuclei tend to undergo spontaneous fission. We therefore predict that
Fm will decay by either or both of these two processes. In fact, it decays by both spontaneous fission and alpha
256
100

emission, in a 97:3 ratio.

 Exercise 6.2.2

Predict the kind of nuclear change each unstable nuclide undergoes when it decays.
a. 32
14
Si

6.2.6 https://chem.libretexts.org/@go/page/13191
b. 43
21
Sc

c. 231
91
Pa

Answer
a. beta decay
b. positron emission or electron capture
c. alpha decay

Radioactive Decay Series


The nuclei of all elements with atomic numbers greater than 83 are unstable. Thus all isotopes of all elements beyond bismuth in
the periodic table are radioactive. Because alpha decay decreases Z by only 2, and positron emission or electron capture decreases Z
by only 1, it is impossible for any nuclide with Z > 85 to decay to a stable daughter nuclide in a single step, except via nuclear
fission. Consequently, radioactive isotopes with Z > 85 usually decay to a daughter nucleus that is radiaoctive, which in turn decays
to a second radioactive daughter nucleus, and so forth, until a stable nucleus finally results. This series of sequential alpha- and
beta-decay reactions is called a radioactive decay series. The most common is the uranium-238 decay series, which produces lead-
206 in a series of 14 sequential alpha- and beta-decay reactions (Figure 6.2.2). Although a radioactive decay series can be written
for almost any isotope with Z > 85, only two others occur naturally: the decay of uranium-235 to lead-207 (in 11 steps) and
thorium-232 to lead-208 (in 10 steps). A fourth series, the decay of neptunium-237 to bismuth-209 in 11 steps, is known to have
occurred on the primitive Earth. With a half-life of “only” 2.14 million years, all the neptunium-237 present when Earth was
formed decayed long ago, and today all the neptunium on Earth is synthetic.

Figure 6.2.1 : A Radioactive Decay Series. Three naturally occurring radioactive decay series are known to occur currently: the
uranium-238 decay series, the decay of uranium-235 to lead-207, and the decay of thorium-232 to lead-208.
Due to these radioactive decay series, small amounts of very unstable isotopes are found in ores that contain uranium or thorium.
These rare, unstable isotopes should have decayed long ago to stable nuclei with a lower atomic number, and they would no longer
be found on Earth. Because they are generated continuously by the decay of uranium or thorium, however, their amounts have
reached a steady state, in which their rate of formation is equal to their rate of decay. In some cases, the abundance of the daughter
isotopes can be used to date a material or identify its origin.

Induced Nuclear Reactions


The discovery of radioactivity in the late 19th century showed that some nuclei spontaneously transform into nuclei with a different
number of protons, thereby producing a different element. When scientists realized that these naturally occurring radioactive
isotopes decayed by emitting subatomic particles, they realized that—in principle—it should be possible to carry out the reverse
reaction, converting a stable nucleus to another more massive nucleus by bombarding it with subatomic particles in a nuclear
transmutation reaction.
The first successful nuclear transmutation reaction was carried out in 1919 by Ernest Rutherford, who showed that α particles
emitted by radium could react with nitrogen nuclei to form oxygen nuclei. As shown in the following equation, a proton is emitted
in the process:

6.2.7 https://chem.libretexts.org/@go/page/13191
4 14 17 1
α+ N → O+ p (6.2.19)
2 7 8 1

Rutherford’s nuclear transmutation experiments led to the discovery of the neutron. He found that bombarding the nucleus of a
light target element with an α particle usually converted the target nucleus to a product that had an atomic number higher by 1 and a
mass number higher by 3 than the target nucleus. Such behavior is consistent with the emission of a proton after reaction with the α
particle. Very light targets such as Li, Be, and B reacted differently, however, emitting a new kind of highly penetrating radiation
rather than a proton. Because neither a magnetic field nor an electrical field could deflect these high-energy particles, Rutherford
concluded that they were electrically neutral. Other observations suggested that the mass of the neutral particle was similar to the
mass of the proton. In 1932, James Chadwick (Nobel Prize in Physics, 1935), who was a student of Rutherford’s at the time, named
these neutral particles neutrons and proposed that they were fundamental building blocks of the atom. The reaction that Chadwick
initially used to explain the production of neutrons was as follows:
4 9 12 1
α+ Be → C+ n (6.2.20)
2 4 6 0

Because α particles and atomic nuclei are both positively charged, electrostatic forces cause them to repel each other. Only α
particles with very high kinetic energy can overcome this repulsion and collide with a nucleus (Figure 6.2.3). Neutrons have no
electrical charge, however, so they are not repelled by the nucleus. Hence bombardment with neutrons is a much easier way to
prepare new isotopes of the lighter elements. In fact, carbon-14 is formed naturally in the atmosphere by bombarding nitrogen-14
with neutrons generated by cosmic rays:
1 14 14 1
n+ N → C+ p (6.2.21)
0 7 6 1

Figure 6.2.3 : A Nuclear Transmutation Reaction. Bombarding a target of one element with high-energy nuclei or subatomic
particles can create new elements. Electrostatic repulsions normally prevent a positively charged particle from colliding and
reacting with a positively charged nucleus. If the positively charged particle is moving at a very high speed, however, its kinetic
energy may be great enough to overcome the electrostatic repulsions, and it may collide with the target nucleus. Such collisions can
result in a nuclear transmutation reaction.

 Example 6.2.3

In 1933, Frédéric Joliot and Iréne Joliot-Curie (daughter of Marie and Pierre Curie) prepared the first artificial radioactive
isotope by bombarding aluminum-27 with α particles. For each 27Al that reacted, one neutron was released. Identify the
product nuclide and write a balanced nuclear equation for this transmutation reaction.
Given: reactants in a nuclear transmutation reaction
Asked for: product nuclide and balanced nuclear equation
Strategy:
A Based on the reactants and one product, identify the other product of the reaction. Use conservation of mass and charge to
determine the values of Z and A of the product nuclide and thus its identity.
B Write the balanced nuclear equation for the reaction.
Solution
A Bombarding an element with α particles usually produces an element with an atomic number that is 2 greater than the atomic
number of the target nucleus. Thus we expect that aluminum (Z = 13) will be converted to phosphorus (Z = 15). With one
neutron released, conservation of mass requires that the mass number of the other product be 3 greater than the mass number of
the target. In this case, the mass number of the target is 27, so the mass number of the product will be 30. The second product
is therefore phosphorus-30, P.30
15

6.2.8 https://chem.libretexts.org/@go/page/13191
B The balanced nuclear equation for the reaction is as follows:
27 4 30 1
Al + α → P+ n (6.2.22)
13 2 15 0

 Exercise 6.2.1

Because all isotopes of technetium are radioactive and have short half-lives, it does not exist in nature. Technetium can,
however, be prepared by nuclear transmutation reactions. For example, bombarding a molybdenum-96 target with deuterium
nuclei ( H) produces technetium-97. Identify the other product of the reaction and write a balanced nuclear equation for this
2
1

transmutation reaction.
Answer
neutron, 1
0
n ; 96
42
Mo +
2
1
H →
97
43
Tc +
1
0
n

We noted earlier in this section that very heavy nuclides, corresponding to Z ≥ 104, tend to decay by spontaneous fission. Nuclides
with slightly lower values of Z, such as the isotopes of uranium (Z = 92) and plutonium (Z = 94), do not undergo spontaneous
fission at any significant rate. Some isotopes of these elements, however, such as U and Pu undergo induced nuclear fission 235
92
239
94

when they are bombarded with relatively low-energy neutrons, as shown in the following equation for uranium-235 and in Figure
6.2.3:

235 1 236 141 92 1


U+ n→ U → Ba + Kr + 3 n (6.2.23)
92 0 92 56 36 0

Figure 6.2.4 : Neutron-Induced Nuclear Fission. Collision of a relatively slow-moving neutron with a fissile nucleus can split it into
two smaller nuclei with the same or different masses. Neutrons are also released in the process, along with a great deal of energy.
Any isotope that can undergo a nuclear fission reaction when bombarded with neutrons is called a fissile isotope.
During nuclear fission, the nucleus usually divides asymmetrically rather than into two equal parts, as shown in Figure 6.2.4.
Moreover, every fission event of a given nuclide does not give the same products; more than 50 different fission modes have been
identified for uranium-235, for example. Consequently, nuclear fission of a fissile nuclide can never be described by a single
equation. Instead, as shown in Figure 6.2.5, a distribution of many pairs of fission products with different yields is obtained, but the
mass ratio of each pair of fission products produced by a single fission event is always roughly 3:2.

Figure 6.2.5 : Mass Distribution of Nuclear Fission Products of 235U. Nuclear fission usually produces a range of products with
different masses and yields, although the mass ratio of each pair of fission products from a fission event is approximately 3:2. As
shown in this plot, more than 50 different fission products are known for235U. Data source: T. R. England and B. F. Rider, Los
Alamos National Laboratory, LA-UR-94-3106, ENDF-349 (1993).

6.2.9 https://chem.libretexts.org/@go/page/13191
Synthesis of Transuranium Elements
Uranium (Z = 92) is the heaviest naturally occurring element. Consequently, all the elements with Z > 92, the transuranium
elements, are artificial and have been prepared by bombarding suitable target nuclei with smaller particles. The first of the
transuranium elements to be prepared was neptunium (Z = 93), which was synthesized in 1940 by bombarding a 238U target with
neutrons. As shown in Equation 20.21, this reaction occurs in two steps. Initially, a neutron combines with a 238U nucleus to form
239U, which is unstable and undergoes beta decay to produce 239Np:

238 1 239 239 0


U+ n→ U → Np + β
92 0 92 93 −1

Subsequent beta decay of 239Np produces the second transuranium element, plutonium (Z = 94):
239 239 0
Np → Pu + β
93 94 −1

Bombarding the target with more massive nuclei creates elements that have atomic numbers significantly greater than that of the
target nucleus (Table 6.2.2). Such techniques have resulted in the creation of the superheavy elements 114 and 116, both of which
lie in or near the “island of stability."
Table 6.2.2 : Some Reactions Used to Synthesize Transuranium Elements
239 4 242 1
Pu + α → Cm + n
94 2 96 0

239 4 241 1 1
Pu + α → Am + p+ n
94 2 95 1 0

242 4 243 1 1
Cm + α → Bk + p + 20 n
96 2 97 1

253 4 256 1
Es + α → Md + n
99 2 101 0

238 12 246 1
U + C → Cf + 40 n
92 6 98

252 10 256 1
Cf + B → Lr + 60 n
98 5 103

A device called a particle accelerator is used to accelerate positively charged particles to the speeds needed to overcome the
electrostatic repulsions between them and the target nuclei by using electrical and magnetic fields. Operationally, the simplest
particle accelerator is the linear accelerator (Figure 6.2.6), in which a beam of particles is injected at one end of a long evacuated
tube. Rapid alternation of the polarity of the electrodes along the tube causes the particles to be alternately accelerated toward a
region of opposite charge and repelled by a region with the same charge, resulting in a tremendous acceleration as the particle
travels down the tube. A modern linear accelerator such as the Stanford Linear Accelerator (SLAC) at Stanford University is about
2 miles long.

Figure 6.2.6 : A Linear Particle Accelerator. (a) An aerial view of the SLAC, the longest linear particle accelerator in the world; the
overall length of the tunnel is 2 miles. (b) Rapidly reversing the polarity of the electrodes in the tube causes the charged particles to
be alternately attracted as they enter one section of the tube and repelled as they leave that section. As a result, the particles are
continuously accelerated along the length of the tube.
To achieve the same outcome in less space, a particle accelerator called a cyclotron forces the charged particles to travel in a
circular path rather than a linear one. The particles are injected into the center of a ring and accelerated by rapidly alternating the
polarity of two large D-shaped electrodes above and below the ring, which accelerates the particles outward along a spiral path
toward the target.
The length of a linear accelerator and the size of the D-shaped electrodes in a cyclotron severely limit the kinetic energy that
particles can attain in these devices. These limitations can be overcome by using a synchrotron, a hybrid of the two designs. A
synchrotron contains an evacuated tube similar to that of a linear accelerator, but the tube is circular and can be more than a mile in
diameter. Charged particles are accelerated around the circle by a series of magnets whose polarities rapidly alternate.

6.2.10 https://chem.libretexts.org/@go/page/13191
Summary
In nuclear decay reactions (or radioactive decay), the parent nucleus is converted to a more stable daughter nucleus. Nuclei with
too many neutrons decay by converting a neutron to a proton, whereas nuclei with too few neutrons decay by converting a proton
to a neutron. Very heavy nuclei (with A ≥ 200 and Z > 83) are unstable and tend to decay by emitting an α particle. When an
unstable nuclide undergoes radioactive decay, the total number of nucleons is conserved, as is the total positive charge. Six
different kinds of nuclear decay reactions are known. Alpha decay results in the emission of an α particle, α , and produces a
4
2

daughter nucleus with a mass number that is lower by 4 and an atomic number that is lower by 2 than the parent nucleus. Beta
decay converts a neutron to a proton and emits a high-energy electron, producing a daughter nucleus with the same mass number as
the parent and an atomic number that is higher by 1. Positron emission is the opposite of beta decay and converts a proton to a
neutron plus a positron. Positron emission does not change the mass number of the nucleus, but the atomic number of the daughter
nucleus is lower by 1 than the parent. In electron capture (EC), an electron in an inner shell reacts with a proton to produce a
neutron, with emission of an x-ray. The mass number does not change, but the atomic number of the daughter is lower by 1 than the
parent. In gamma emission, a daughter nucleus in a nuclear excited state undergoes a transition to a lower-energy state by emitting
a γ ray. Very heavy nuclei with high neutron-to-proton ratios can undergo spontaneous fission, in which the nucleus breaks into
two pieces that can have different atomic numbers and atomic masses with the release of neutrons. Many very heavy nuclei decay
via a radioactive decay series—a succession of some combination of alpha- and beta-decay reactions. In nuclear transmutation
reactions, a target nucleus is bombarded with energetic subatomic particles to give a product nucleus that is more massive than the
original. All transuranium elements—elements with Z > 92—are artificial and must be prepared by nuclear transmutation
reactions. These reactions are carried out in particle accelerators such as linear accelerators, cyclotrons, and synchrotrons.

Key Takeaway
Nuclear decay reactions occur spontaneously under all conditions and produce more stable daughter nuclei, whereas nuclear
transmutation reactions are induced and form a product nucleus that is more massive than the starting material.

Key Equations
alpha decay
Equation 20.1: A
Z
X →
A−4

Z−2

X +
4
2
α

beta decay
Equation 20.4: A
Z
X →
A
Z+1
X +
′ 0
−1
β

positron emission
Equation 20.7: A
Z
X →
A
Z−1
X +
′ 0
+1
β

electron capture
Equation 20.10: A
Z
X+
0
−1
e →
A
Z−1

X + x-ray

gamma emission
Equation 20.15: A
Z
X* →
A
Z
X+
0
0
γ

6.2: Nuclear Reactions is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
24.3: Nuclear Reactions by Anonymous is licensed CC BY-NC-SA 3.0.

6.2.11 https://chem.libretexts.org/@go/page/13191
6.3: Nuclear Radiation
 Learning Objectives
Write and balance nuclear equations
To know the different kinds of radioactive decay.
To balance a nuclear reaction.

Nuclear chemistry is the study of reactions that involve changes in nuclear structure. The chapter on atoms, molecules, and ions
introduced the basic idea of nuclear structure, that the nucleus of an atom is composed of protons and, with the exception of H , 1
1

neutrons. Recall that the number of protons in the nucleus is called the atomic number (Z ) of the element, and the sum of the
number of protons and the number of neutrons is the mass number (A ). Atoms with the same atomic number but different mass
numbers are isotopes of the same element. When referring to a single type of nucleus, we often use the term nuclide and identify it
by the notation:
A
X (6.3.1)
Z

where
X is the symbol for the element,
A is the mass number, and
Z is the atomic number.
Often a nuclide is referenced by the name of the element followed by a hyphen and the mass number. For example, 14
6
C is called
“carbon-14.”
Protons and neutrons, collectively called nucleons, are packed together tightly in a nucleus. With a radius of about 10−15 meters, a
nucleus is quite small compared to the radius of the entire atom, which is about 10−10 meters. Nuclei are extremely dense compared
to bulk matter, averaging 1.8 × 10 grams per cubic centimeter. For example, water has a density of 1 gram per cubic centimeter,
14

and iridium, one of the densest elements known, has a density of 22.6 g/cm3. If the earth’s density were equal to the average
nuclear density, the earth’s radius would be only about 200 meters (earth’s actual radius is approximately 6.4 × 10 meters, 30,000
6

times larger).
Changes of nuclei that result in changes in their atomic numbers, mass numbers, or energy states are nuclear reactions. To describe
a nuclear reaction, we use an equation that identifies the nuclides involved in the reaction, their mass numbers and atomic numbers,
and the other particles involved in the reaction.

Nuclear Equations
A balanced chemical reaction equation reflects the fact that during a chemical reaction, bonds break and form, and atoms are
rearranged, but the total numbers of atoms of each element are conserved and do not change. A balanced nuclear reaction equation
indicates that there is a rearrangement during a nuclear reaction, but of subatomic particles rather than atoms. Nuclear reactions
also follow conservation laws, and they are balanced in two ways:
1. The sum of the mass numbers of the reactants equals the sum of the mass numbers of the products.
2. The sum of the charges of the reactants equals the sum of the charges of the products.
If the atomic number and the mass number of all but one of the particles in a nuclear reaction are known, we can identify the
particle by balancing the reaction. For instance, we could determine that O is a product of the nuclear reaction of N and He if
17
8
14
7
4
2

we knew that a proton, H , was one of the two products. Example 6.3.1 shows how we can identify a nuclide by balancing the
1
1

nuclear reaction.

 Example 6.3.1: Balancing Equations for Nuclear Reactions

The reaction of an α particle with magnesium-25 ( 25


12
Mg ) produces a proton and a nuclide of another element. Identify the new
nuclide produced.

6.3.1 https://chem.libretexts.org/@go/page/13192
Solution
The nuclear reaction can be written as:
25 4 1 A
12
Mg + 2 He → 1
H+ Z
X

where
A is the mass number and
Z is the atomic number of the new nuclide, X .
Because the sum of the mass numbers of the reactants must equal the sum of the mass numbers of the products:

25 + 4 = A + 1

so
A = 28

Similarly, the charges must balance, so:

12 + 2 = Z + 1

so

Z = 13

Check the periodic table: The element with nuclear charge = +13 is aluminum. Thus, the product is 28
13
Al .

 Exercise 6.3.1

The nuclide I combines with an electron and produces a new nucleus and no other massive particles. What is the equation
125
53

for this reaction?

Answer
125 0 125
53
I+ e → 52
Te
−1

The two general kinds of nuclear reactions are nuclear decay reactions and nuclear transmutation reactions. In a nuclear decay
reaction, also called radioactive decay, an unstable nucleus emits radiation and is transformed into the nucleus of one or more other
elements. The resulting daughter nuclei have a lower mass and are lower in energy (more stable) than the parent nucleus that
decayed. In contrast, in a nuclear transmutation reaction, a nucleus reacts with a subatomic particle or another nucleus to form a
product nucleus that is more massive than the starting material. As we shall see, nuclear decay reactions occur spontaneously under
all conditions, but nuclear transmutation reactions occur only under very special conditions, such as the collision of a beam of
highly energetic particles with a target nucleus or in the interior of stars. We begin this section by considering the different classes
of radioactive nuclei, along with their characteristic nuclear decay reactions and the radiation they emit.

Nuclear decay reactions occur spontaneously under all conditions, whereas nuclear
transmutation reactions are induced.

Nuclear Decay Reactions


Just as we use the number and type of atoms present to balance a chemical equation, we can use the number and type of nucleons
present to write a balanced nuclear equation for a nuclear decay reaction. This procedure also allows us to predict the identity of
either the parent or the daughter nucleus if the identity of only one is known. Regardless of the mode of decay, the total number of
nucleons is conserved in all nuclear reactions.
To describe nuclear decay reactions, chemists have extended the X notation for nuclides to include radioactive emissions. Table
A
Z

6.3.1 lists the name and symbol for each type of emitted radiation. The most notable addition is the positron, a particle that has the

same mass as an electron but a positive charge rather than a negative charge.
Table 6.3.1 : Nuclear Decay Emissions and Their Symbols

6.3.2 https://chem.libretexts.org/@go/page/13192
Identity Symbol Charge Mass (amu)

helium nucleus 4
2
α +2 4.001506

electron 0
−1
β or β −
−1 0.000549

photon 0
0
γ — —

neutron 1
0
n 0 1.008665

proton 1
1
p +1 1.007276

positron 0
+1
β or β +
+1 0.000549

Like the notation used to indicate isotopes, the upper left superscript in the symbol for a particle gives the mass number, which is
the total number of protons and neutrons. For a proton or a neutron, A = 1. Because neither an electron nor a positron contains
protons or neutrons, its mass number is 0. The numbers should not be taken literally, however, as meaning that these particles have
zero mass; ejection of a beta particle (an electron) simply has a negligible effect on the mass of a nucleus.
Similarly, the lower left subscript gives the charge of the particle. Because protons carry a positive charge, Z = +1 for a proton. In
contrast, a neutron contains no protons and is electrically neutral, so Z = 0. In the case of an electron, Z = −1, and for a positron, Z =
+1. Because γ rays are high-energy photons, both A and Z are 0. In some cases, two different symbols are used for particles that are
identical but produced in different ways. For example, the symbol e, which is usually simplified to e−, represents a free electron
0
−1

or an electron associated with an atom, whereas the symbol β, which is often simplified to β−, denotes an electron that originates
0
−1

from within the nucleus, which is a β particle. Similarly, He refers to the nucleus of a helium atom, and α denotes an identical
4
2
2+ 4
2

particle that has been ejected from a heavier nucleus.


There are six fundamentally different kinds of nuclear decay reactions, and each releases a different kind of particle or energy. The
essential features of each reaction are shown in Figure 6.3.1. The most common are alpha and beta decay and gamma emission, but
the others are essential to an understanding of nuclear decay reactions.

6.3.3 https://chem.libretexts.org/@go/page/13192
Figure 6.3.1 : Common Modes of Nuclear Decay
The different types of decay are alpha, beta, positron emission, electron capture, gamma emission, and spontaneous fission.

Alpha α Decay
Many nuclei with mass numbers greater than 200 undergo alpha (α) decay, which results in the emission of a helium-4 nucleus as
an alpha (α) particle, α . The general reaction is as follows:
4
2

A A−4 ′ 4
X → X + α (6.3.2)
Z Z−2 2
parent daughter alpha

particle

The daughter nuclide contains two fewer protons and two fewer neutrons than the parent. Thus α-particle emission produces a
daughter nucleus with a mass number A − 4 and a nuclear charge Z − 2 compared to the parent nucleus. Radium-226, for example,
undergoes alpha decay to form radon-222:
226 222 4
Ra → Rn + α (6.3.3)
88 86 2

Because nucleons are conserved in this and all other nuclear reactions, the sum of the mass numbers of the products, 222 + 4 = 226,
equals the mass number of the parent. Similarly, the sum of the atomic numbers of the products, 86 + 2 = 88, equals the atomic
number of the parent. Thus the nuclear equation is balanced.

6.3.4 https://chem.libretexts.org/@go/page/13192
Just as the total number of atoms is conserved in a chemical reaction, the total number of
nucleons is conserved in a nuclear reaction.

Beta β −
Decay
Nuclei that contain too many neutrons often undergo beta (β) decay, in which a neutron is converted to a proton and a high-energy
electron that is ejected from the nucleus as a β particle:
1 1 0
n → p + β (6.3.4)
0 1 −1
unstable proton beta particle

neutron in retained emitted by

nucleus by nucleus
nucleus

The general reaction for beta decay is therefore


A A ′ 0
X → X + β (6.3.5)
Z Z+1 −1
parent daughter beta particle

Although beta decay does not change the mass number of the nucleus, it does result in an increase of +1 in the atomic number
because of the addition of a proton in the daughter nucleus. Thus beta decay decreases the neutron-to-proton ratio, moving the
nucleus toward the band of stable nuclei. For example, carbon-14 undergoes beta decay to form nitrogen-14:
14 14 0
C → N+ β
6 7 −1

Once again, the number of nucleons is conserved, and the charges are balanced. The parent and the daughter nuclei have the same
mass number, 14, and the sum of the atomic numbers of the products is 6, which is the same as the atomic number of the carbon-14
parent.

Positron β +
Emission
Because a positron has the same mass as an electron but opposite charge, positron emission is the opposite of beta decay. Thus
positron emission is characteristic of neutron-poor nuclei, which decay by transforming a proton to a neutron and emitting a high-
energy positron:
1 + 1 0 +
p → n+ β (6.3.6)
1 0 +1

The general reaction for positron emission is therefore


A A ′ 0 +
X → X + β
Z Z−1 +1
parent daughter positron

Like beta decay, positron emission does not change the mass number of the nucleus. In this case, however, the atomic number of
the daughter nucleus is lower by 1 than that of the parent. Thus the neutron-to-proton ratio has increased, again moving the nucleus
closer to the band of stable nuclei. For example, carbon-11 undergoes positron emission to form boron-11:
11 11 0 +
C → B+ β
6 5 +1

Nucleons are conserved, and the charges balance. The mass number, 11, does not change, and the sum of the atomic numbers of the
products is 6, the same as the atomic number of the parent carbon-11 nuclide.

Electron Capture
A neutron-poor nucleus can decay by either positron emission or electron capture (EC), in which an electron in an inner shell
reacts with a proton to produce a neutron:
1 0 1
p+ e → n (6.3.7)
1 −1 0

When a second electron moves from an outer shell to take the place of the lower-energy electron that was absorbed by the nucleus,
an x-ray is emitted. The overall reaction for electron capture is thus
A 0 A ′
X + e → X + x-ray
Z −1 Z−1
parent electron daughter

6.3.5 https://chem.libretexts.org/@go/page/13192
Electron capture does not change the mass number of the nucleus because both the proton that is lost and the neutron that is formed
have a mass number of 1. As with positron emission, however, the atomic number of the daughter nucleus is lower by 1 than that of
the parent. Once again, the neutron-to-proton ratio has increased, moving the nucleus toward the band of stable nuclei. For
example, iron-55 decays by electron capture to form manganese-55, which is often written as follows:
EC
55 55
Fe → Mn + x-ray
26 25

The atomic numbers of the parent and daughter nuclides differ in Equation 20.2.11, although the mass numbers are the same. To
write a balanced nuclear equation for this reaction, we must explicitly include the captured electron in the equation:
55 0 55
Fe + e → Mn + x-ray
26 −1 25

Both positron emission and electron capture are usually observed for nuclides with low neutron-to-proton ratios, but the decay rates
for the two processes can be very different.

Gamma γ Emission
Many nuclear decay reactions produce daughter nuclei that are in a nuclear excited state, which is similar to an atom in which an
electron has been excited to a higher-energy orbital to give an electronic excited state. Just as an electron in an electronic excited
state emits energy in the form of a photon when it returns to the ground state, a nucleus in an excited state releases energy in the
form of a photon when it returns to the ground state. These high-energy photons are γ rays. Gamma (γ) emission can occur
virtually instantaneously, as it does in the alpha decay of uranium-238 to thorium-234, where the asterisk denotes an excited state:
relaxation
238 234 4 234 0
U → Th* + α −−−−−−→ Th + γ
92 90 2 90 0
excited

nuclear

state

If we disregard the decay event that created the excited nucleus, then
234 234 0
Th* → Th + γ
88 88 0

or more generally,
A A 0
X* → X+ γ
Z Z 0

Gamma emission can also occur after a significant delay. For example, technetium-99m has a half-life of about 6 hours before
emitting a γ ray to form technetium-99 (the m is for metastable). Because γ rays are energy, their emission does not affect either the
mass number or the atomic number of the daughter nuclide. Gamma-ray emission is therefore the only kind of radiation that does
not necessarily involve the conversion of one element to another, although it is almost always observed in conjunction with some
other nuclear decay reaction.

Spontaneous Fission
Only very massive nuclei with high neutron-to-proton ratios can undergo spontaneous fission, in which the nucleus breaks into
two pieces that have different atomic numbers and atomic masses. This process is most important for the transactinide elements,
with Z ≥ 104. Spontaneous fission is invariably accompanied by the release of large amounts of energy, and it is usually
accompanied by the emission of several neutrons as well. An example is the spontaneous fission of Cf , which gives a
254
98

distribution of fission products; one possible set of products is shown in the following equation:
254 118 132 1
Cf → Pd + Te + 4 n
98 46 52 0

Once again, the number of nucleons is conserved. Thus the sum of the mass numbers of the products (118 + 132 + 4 = 254) equals
the mass number of the reactant. Similarly, the sum of the atomic numbers of the products [46 + 52 + (4 × 0) = 98] is the same as
the atomic number of the parent nuclide.

 Example 6.3.2
Write a balanced nuclear equation to describe each reaction.
a. the beta decay of 35
16
S

6.3.6 https://chem.libretexts.org/@go/page/13192
b. the decay of 201
80
Hg by electron capture
c. the decay of 30
15
P by positron emission
Given: radioactive nuclide and mode of decay
Asked for: balanced nuclear equation

Strategy:
A Identify the reactants and the products from the information given.
B Use the values of A and Z to identify any missing components needed to balance the equation.

Solution
a.
A We know the identities of the reactant and one of the products (a β particle). We can therefore begin by writing an
equation that shows the reactant and one of the products and indicates the unknown product as X: A
Z

35 A 0
S→ X+ β
16 Z −1

B Because both protons and neutrons must be conserved in a nuclear reaction, the unknown product must have a mass
number of A = 35 − 0 = 35 and an atomic number of Z = 16 − (−1) = 17. The element with Z = 17 is chlorine, so the
balanced nuclear equation is as follows:
35 35 0
S→ Cl + β
16 17 −1

b.
A We know the identities of both reactants: 201
80
Hg and an inner electron, 0
−1
. The reaction is as follows:
e

201 0 A
Hg + e → X
80 −1 Z

B Both protons and neutrons are conserved, so the mass number of the product must be A = 201 + 0 = 201, and the
atomic number of the product must be Z = 80 + (−1) = 79, which corresponds to the element gold. The balanced nuclear
equation is thus
201 0 201
Hg + e → Au
80 −1 79

c.
A As in part (a), we are given the identities of the reactant and one of the products—in this case, a positron. The
unbalanced nuclear equation is therefore
30 A 0
P → X+ β
15 Z +1

B The mass number of the second product is A = 30 − 0 = 30, and its atomic number is Z = 15 − 1 = 14, which
corresponds to silicon. The balanced nuclear equation for the reaction is as follows:
30 30 0
P → Si + β
15 14 +1

 Exercise 6.3.2

Write a balanced nuclear equation to describe each reaction.


a. C by positron emission
11
6

b. the beta decay of molybdenum-99


c. the emission of an α particle followed by gamma emission from 185
74
W

Answer a

6.3.7 https://chem.libretexts.org/@go/page/13192
11 11 0
C → B+ β
6 5 +1

Answer d
99 99m 0
Mo → Tc + β
42 43 −1

Answer c
185 181 4 0
W → Hf + α+ γ
74 72 2 0

 Example 6.3.3

Predict the kind of nuclear change each unstable nuclide undergoes when it decays.
a. 45
22
Ti

b. 242
94
Pu

c. 12
5
B

d. 256
100
Fm

Given: nuclide
Asked for: type of nuclear decay

Strategy:
Based on the neutron-to-proton ratio and the value of Z, predict the type of nuclear decay reaction that will produce a more
stable nuclide.

Solution
a. This nuclide has a neutron-to-proton ratio of only 1.05, which is much less than the requirement for stability for an element
with an atomic number in this range. Nuclei that have low neutron-to-proton ratios decay by converting a proton to a
neutron. The two possibilities are positron emission, which converts a proton to a neutron and a positron, and electron
capture, which converts a proton and a core electron to a neutron. In this case, both are observed, with positron emission
occurring about 86% of the time and electron capture about 14% of the time.
b. Nuclei with Z > 83 are too heavy to be stable and usually undergo alpha decay, which decreases both the mass number and
the atomic number. Thus Pu is expected to decay by alpha emission.
242
94

c. This nuclide has a neutron-to-proton ratio of 1.4, which is very high for a light element. Nuclei with high neutron-to-proton
ratios decay by converting a neutron to a proton and an electron. The electron is emitted as a β particle, and the proton
remains in the nucleus, causing an increase in the atomic number with no change in the mass number. We therefore predict
that B will undergo beta decay.
12
5

d. This is a massive nuclide, with an atomic number of 100 and a mass number much greater than 200. Nuclides with A ≥ 200
tend to decay by alpha emission, and even heavier nuclei tend to undergo spontaneous fission. We therefore predict that
Fm will decay by either or both of these two processes. In fact, it decays by both spontaneous fission and alpha
256
100

emission, in a 97:3 ratio.

 Exercise 6.3.3

Predict the kind of nuclear change each unstable nuclide undergoes when it decays.
a. 32
14
Si

b. 43
21
Sc

c. 231
91
Pa

Answer a
beta decay

6.3.8 https://chem.libretexts.org/@go/page/13192
Answer d
positron emission or electron capture
Answer c
alpha decay

Radioactive Decay Series


The nuclei of all elements with atomic numbers greater than 83 are unstable. Thus all isotopes of all elements beyond bismuth in
the periodic table are radioactive. Because alpha decay decreases Z by only 2, and positron emission or electron capture decreases Z
by only 1, it is impossible for any nuclide with Z > 85 to decay to a stable daughter nuclide in a single step, except via nuclear
fission. Consequently, radioactive isotopes with Z > 85 usually decay to a daughter nucleus that is radiaoctive, which in turn decays
to a second radioactive daughter nucleus, and so forth, until a stable nucleus finally results. This series of sequential alpha- and
beta-decay reactions is called a radioactive decay series. The most common is the uranium-238 decay series, which produces lead-
206 in a series of 14 sequential alpha- and beta-decay reactions (Figure 6.3.2). Although a radioactive decay series can be written
for almost any isotope with Z > 85, only two others occur naturally: the decay of uranium-235 to lead-207 (in 11 steps) and
thorium-232 to lead-208 (in 10 steps). A fourth series, the decay of neptunium-237 to bismuth-209 in 11 steps, is known to have
occurred on the primitive Earth. With a half-life of “only” 2.14 million years, all the neptunium-237 present when Earth was
formed decayed long ago, and today all the neptunium on Earth is synthetic.

Figure 6.3.2 A Radioactive Decay Series. Three naturally occurring radioactive decay series are known to occur currently: the
uranium-238 decay series, the decay of uranium-235 to lead-207, and the decay of thorium-232 to lead-208.
Graph of mass number against number of protons. The purple lines are alpha decay which are linear while the green lines are beta
decay and are horizontal and parallel to the x axis.
Due to these radioactive decay series, small amounts of very unstable isotopes are found in ores that contain uranium or thorium.
These rare, unstable isotopes should have decayed long ago to stable nuclei with a lower atomic number, and they would no longer
be found on Earth. Because they are generated continuously by the decay of uranium or thorium, however, their amounts have
reached a steady state, in which their rate of formation is equal to their rate of decay. In some cases, the abundance of the daughter
isotopes can be used to date a material or identify its origin.

Induced Nuclear Reactions


The discovery of radioactivity in the late 19th century showed that some nuclei spontaneously transform into nuclei with a different
number of protons, thereby producing a different element. When scientists realized that these naturally occurring radioactive
isotopes decayed by emitting subatomic particles, they realized that—in principle—it should be possible to carry out the reverse
reaction, converting a stable nucleus to another more massive nucleus by bombarding it with subatomic particles in a nuclear
transmutation reaction.

6.3.9 https://chem.libretexts.org/@go/page/13192
The first successful nuclear transmutation reaction was carried out in 1919 by Ernest Rutherford, who showed that α particles
emitted by radium could react with nitrogen nuclei to form oxygen nuclei. As shown in the following equation, a proton is emitted
in the process:
4 14 17 1
α+ N → O+ p
2 7 8 1

Rutherford’s nuclear transmutation experiments led to the discovery of the neutron. He found that bombarding the nucleus of a
light target element with an α particle usually converted the target nucleus to a product that had an atomic number higher by 1 and a
mass number higher by 3 than the target nucleus. Such behavior is consistent with the emission of a proton after reaction with the α
particle. Very light targets such as Li, Be, and B reacted differently, however, emitting a new kind of highly penetrating radiation
rather than a proton. Because neither a magnetic field nor an electrical field could deflect these high-energy particles, Rutherford
concluded that they were electrically neutral. Other observations suggested that the mass of the neutral particle was similar to the
mass of the proton. In 1932, James Chadwick (Nobel Prize in Physics, 1935), who was a student of Rutherford’s at the time, named
these neutral particles neutrons and proposed that they were fundamental building blocks of the atom. The reaction that Chadwick
initially used to explain the production of neutrons was as follows:
4 9 12 1
α+ Be → C+ n (6.3.8)
2 4 6 0

Because α particles and atomic nuclei are both positively charged, electrostatic forces cause them to repel each other. Only α
particles with very high kinetic energy can overcome this repulsion and collide with a nucleus (Figure 6.3.3). Neutrons have no
electrical charge, however, so they are not repelled by the nucleus. Hence bombardment with neutrons is a much easier way to
prepare new isotopes of the lighter elements. In fact, carbon-14 is formed naturally in the atmosphere by bombarding nitrogen-14
with neutrons generated by cosmic rays:
1 14 14 1
n+ N → C+ p (6.3.9)
0 7 6 1

Figure 6.3.3 : A Nuclear Transmutation Reaction. Bombarding a target of one element with high-energy nuclei or subatomic
particles can create new elements. Electrostatic repulsions normally prevent a positively charged particle from colliding and
reacting with a positively charged nucleus. If the positively charged particle is moving at a very high speed, however, its kinetic
energy may be great enough to overcome the electrostatic repulsions, and it may collide with the target nucleus. Such collisions can
result in a nuclear transmutation reaction.

 Example 6.3.4

In 1933, Frédéric Joliot and Iréne Joliot-Curie (daughter of Marie and Pierre Curie) prepared the first artificial radioactive
isotope by bombarding aluminum-27 with α particles. For each 27Al that reacted, one neutron was released. Identify the
product nuclide and write a balanced nuclear equation for this transmutation reaction.
Given: reactants in a nuclear transmutation reaction
Asked for: product nuclide and balanced nuclear equation

Strategy:
A Based on the reactants and one product, identify the other product of the reaction. Use conservation of mass and charge to
determine the values of Z and A of the product nuclide and thus its identity.
B Write the balanced nuclear equation for the reaction.

6.3.10 https://chem.libretexts.org/@go/page/13192
Solution
A Bombarding an element with α particles usually produces an element with an atomic number that is 2 greater than the atomic
number of the target nucleus. Thus we expect that aluminum (Z = 13) will be converted to phosphorus (Z = 15). With one
neutron released, conservation of mass requires that the mass number of the other product be 3 greater than the mass number of
the target. In this case, the mass number of the target is 27, so the mass number of the product will be 30. The second product
is therefore phosphorus-30, P. 30
15

B The balanced nuclear equation for the reaction is as follows:


27 4 30 1
Al + α → P+ n
13 2 15 0

 Exercise 6.3.4

Because all isotopes of technetium are radioactive and have short half-lives, it does not exist in nature. Technetium can,
however, be prepared by nuclear transmutation reactions. For example, bombarding a molybdenum-96 target with deuterium
nuclei ( H) produces technetium-97. Identify the other product of the reaction and write a balanced nuclear equation for this
2
1

transmutation reaction.

Answer
neutron, 1
0
n ; 96
42
Mo +
2
1
H →
97
43
Tc +
1
0
n :

We noted earlier in this section that very heavy nuclides, corresponding to Z ≥ 104, tend to decay by spontaneous fission. Nuclides
with slightly lower values of Z, such as the isotopes of uranium (Z = 92) and plutonium (Z = 94), do not undergo spontaneous
fission at any significant rate. Some isotopes of these elements, however, such as U and Pu undergo induced nuclear fission
235
92
239
94

when they are bombarded with relatively low-energy neutrons, as shown in the following equation for uranium-235 and in Figure
6.3.4:

235 1 236 141 92 1


U+ n→ U → Ba + Kr + 30 n (6.3.10)
92 0 92 56 36

Figure 6.3.4 Neutron-Induced Nuclear Fission. Collision of a relatively slow-moving neutron with a fissile nucleus can split it into
two smaller nuclei with the same or different masses. Neutrons are also released in the process, along with a great deal of energy.
Any isotope that can undergo a nuclear fission reaction when bombarded with neutrons is called a fissile isotope.
During nuclear fission, the nucleus usually divides asymmetrically rather than into two equal parts, as shown in Figure 6.3.4.
Moreover, every fission event of a given nuclide does not give the same products; more than 50 different fission modes have been
identified for uranium-235, for example. Consequently, nuclear fission of a fissile nuclide can never be described by a single
equation. Instead, as shown in Figure 6.3.5, a distribution of many pairs of fission products with different yields is obtained, but the
mass ratio of each pair of fission products produced by a single fission event is always roughly 3:2.

6.3.11 https://chem.libretexts.org/@go/page/13192
Figure 6.3.5 : Mass Distribution of Nuclear Fission Products of 235U. Nuclear fission usually produces a range of products with
different masses and yields, although the mass ratio of each pair of fission products from a fission event is approximately 3:2. As
shown in this plot, more than 50 different fission products are known for235U. Data source: T. R. England and B. F. Rider, Los
Alamos National Laboratory, LA-UR-94-3106, ENDF-349 (1993).

Synthesis of Transuranium Elements


Uranium (Z = 92) is the heaviest naturally occurring element. Consequently, all the elements with Z > 92, the transuranium
elements, are artificial and have been prepared by bombarding suitable target nuclei with smaller particles. The first of the
transuranium elements to be prepared was neptunium (Z = 93), which was synthesized in 1940 by bombarding a 238U target with
neutrons. As shown in Equation 20.21, this reaction occurs in two steps. Initially, a neutron combines with a 238U nucleus to form
239
U, which is unstable and undergoes beta decay to produce 239Np:
238 1 239 239 0
U+ n→ U → Np + β
92 0 92 93 −1

Subsequent beta decay of 239Np produces the second transuranium element, plutonium (Z = 94):
239 239 0
Np → Pu + β
93 94 −1

Bombarding the target with more massive nuclei creates elements that have atomic numbers significantly greater than that of the
target nucleus (Table 6.3.2). Such techniques have resulted in the creation of the superheavy elements 114 and 116, both of which
lie in or near the “island of stability."
Table 6.3.2 : Some Reactions Used to Synthesize Transuranium Elements
239 4 242 1
Pu + α → Cm + n
94 2 96 0

239 4 241 1 1
Pu + α → Am + p+ n
94 2 95 1 0

242 4 243 1 1
Cm + α → Bk + p+2 n
96 2 97 1 0

253 4 256 1
Es + α → Md + n
99 2 101 0

238 12 246 1
U + C → Cf + 4 n
92 6 98 0

252 10 256 1
Cf + B → Lr + 6 n
98 5 103 0

A device called a particle accelerator is used to accelerate positively charged particles to the speeds needed to overcome the
electrostatic repulsions between them and the target nuclei by using electrical and magnetic fields. Operationally, the simplest
particle accelerator is the linear accelerator (Figure 6.3.6), in which a beam of particles is injected at one end of a long evacuated
tube. Rapid alternation of the polarity of the electrodes along the tube causes the particles to be alternately accelerated toward a
region of opposite charge and repelled by a region with the same charge, resulting in a tremendous acceleration as the particle
travels down the tube. A modern linear accelerator such as the Stanford Linear Accelerator (SLAC) at Stanford University is about
2 miles long.

6.3.12 https://chem.libretexts.org/@go/page/13192
Figure 6.3.6 : A Linear Particle Accelerator. (a) An aerial view of the SLAC, the longest linear particle accelerator in the world; the
overall length of the tunnel is 2 miles. (b) Rapidly reversing the polarity of the electrodes in the tube causes the charged particles to
be alternately attracted as they enter one section of the tube and repelled as they leave that section. As a result, the particles are
continuously accelerated along the length of the tube.
To achieve the same outcome in less space, a particle accelerator called a cyclotron forces the charged particles to travel in a
circular path rather than a linear one. The particles are injected into the center of a ring and accelerated by rapidly alternating the
polarity of two large D-shaped electrodes above and below the ring, which accelerates the particles outward along a spiral path
toward the target.
The length of a linear accelerator and the size of the D-shaped electrodes in a cyclotron severely limit the kinetic energy that
particles can attain in these devices. These limitations can be overcome by using a synchrotron, a hybrid of the two designs. A
synchrotron contains an evacuated tube similar to that of a linear accelerator, but the tube is circular and can be more than a mile in
diameter. Charged particles are accelerated around the circle by a series of magnets whose polarities rapidly alternate.

Summary and Key Takeaway


Nuclear decay reactions occur spontaneously under all conditions and produce more stable daughter nuclei, whereas nuclear
transmutation reactions are induced and form a product nucleus that is more massive than the starting material.
In nuclear decay reactions (or radioactive decay), the parent nucleus is converted to a more stable daughter nucleus. Nuclei with
too many neutrons decay by converting a neutron to a proton, whereas nuclei with too few neutrons decay by converting a proton
to a neutron. Very heavy nuclei (with A ≥ 200 and Z > 83) are unstable and tend to decay by emitting an α particle. When an
unstable nuclide undergoes radioactive decay, the total number of nucleons is conserved, as is the total positive charge. Six
different kinds of nuclear decay reactions are known. Alpha decay results in the emission of an α particle, α , and produces a 4
2

daughter nucleus with a mass number that is lower by 4 and an atomic number that is lower by 2 than the parent nucleus. Beta
decay converts a neutron to a proton and emits a high-energy electron, producing a daughter nucleus with the same mass number as
the parent and an atomic number that is higher by 1. Positron emission is the opposite of beta decay and converts a proton to a
neutron plus a positron. Positron emission does not change the mass number of the nucleus, but the atomic number of the daughter
nucleus is lower by 1 than the parent. In electron capture (EC), an electron in an inner shell reacts with a proton to produce a
neutron, with emission of an x-ray. The mass number does not change, but the atomic number of the daughter is lower by 1 than the
parent. In gamma emission, a daughter nucleus in a nuclear excited state undergoes a transition to a lower-energy state by emitting
a γ ray. Very heavy nuclei with high neutron-to-proton ratios can undergo spontaneous fission, in which the nucleus breaks into
two pieces that can have different atomic numbers and atomic masses with the release of neutrons. Many very heavy nuclei decay
via a radioactive decay series—a succession of some combination of alpha- and beta-decay reactions. In nuclear transmutation
reactions, a target nucleus is bombarded with energetic subatomic particles to give a product nucleus that is more massive than the
original. All transuranium elements—elements with Z > 92—are artificial and must be prepared by nuclear transmutation
reactions. These reactions are carried out in particle accelerators such as linear accelerators, cyclotrons, and synchrotrons.

Key Equations
alpha decay
A A−4 ′ 4
X → X + α
Z Z−2 2

beta decay

6.3.13 https://chem.libretexts.org/@go/page/13192
A A ′ 0
X → X + β
Z Z+1 −1

positron emission
A A ′ 0
X → X + β
Z Z−1 +1

electron capture
A 0 A ′
X+ e → X + x-ray
Z −1 Z−1

gamma emission
A A 0
X* → X+ γ
Z Z 0

6.3: Nuclear Radiation is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
21.1: Radioactivity is licensed CC BY-NC-SA 3.0.

6.3.14 https://chem.libretexts.org/@go/page/13192
6.4: Rates of Radioactive Decay
Learning Objectives
To know how to use half-lives to describe the rates of first-order reactions

The rate of radioactive decay is an intrinsic property of each radioactive isotope that is independent of the chemical and physical
form of the radioactive isotope. The rate is also independent of temperature. In this section, we will describe radioactive decay rates
and how half-lives can be used to monitor radioactive decay processes.

A Quick Primer from Chemical Kinetics


As discussed previously, the half-life of a first-order reaction is independent of the concentration of the reactants. This becomes
evident when we rearrange the integrated rate law for a first-order reaction to produce the following equation:
[A]0
ln = kt (6.4.1)
[A]

Substituting [A]0/2 for [A] and t1/2 for t (to indicate a half-life) into Equation 6.4.1 gives
[A]0
ln = ln 2 = kt1/2 (6.4.2)
[A]0 /2

Substituting ln 2 ≈ 0.693 into the equation results in the expression for the half-life of a first-order reaction:
0.693
t1/2 = (6.4.3)
k

Thus, for a first-order reaction, each successive half-life is the same length of time. Since radioactivity follows first order decay
kinetics, these equations are also applicable for this type of nuclear reactions.

In any sample of a given radioactive substance, the number of atoms of the radioactive isotope must decrease with time as their
nuclei decay to nuclei of a more stable isotope. Using N to represent the number of atoms of the radioactive isotope, we can define
the rate of decay of the sample, which is also called its activity (A ) as the decrease in the number of the radioisotope’s nuclei per
unit time:
ΔN
A =− (6.4.4)
Δt

Activity is usually measured in disintegrations per second (dps) or disintegrations per minute (dpm).
The activity of a sample is directly proportional to the number of atoms of the radioactive isotope in the sample:

A = kN (6.4.5)

−1 −1
Here, the symbol k is the radioactive decay constant, which has units of inverse time (e.g., s , yr ) and a characteristic value for
each radioactive isotope. If we combine Equation 6.4.4 and Equation 6.4.5, we obtain the relationship between the number of
decays per unit time and the number of atoms of the isotope in a sample:
ΔN
− = kN (6.4.6)
Δt

Equation 6.4.6 is the same as the equation for the reaction rate of a first-order reaction, except that it uses numbers of atoms instead
of concentrations. In fact, radioactive decay is a first-order process and can be described in terms of either the differential rate law
(Equation 6.4.6) or the integrated rate law:
−kt
N = N0 e (6.4.7)

N
ln = −kt (6.4.8)
N0

6.4.1 https://chem.libretexts.org/@go/page/83322
Because radioactive decay is a first-order process, the time required for half of the nuclei in any sample of a radioactive isotope to
decay is a constant, called the half-life of the isotope. The half-life tells us how radioactive an isotope is (the number of decays per
unit time); thus it is the most commonly cited property of any radioisotope. For a given number of atoms, isotopes with shorter
half-lives decay more rapidly, undergoing a greater number of radioactive decays per unit time than do isotopes with longer half-
lives. The half-lives of several isotopes are listed in Table 6.4.1, along with some of their applications.
Table 6.4.1 : Half-Lives and Applications of Some Radioactive Isotopes
Radioactive Isotope Half-Life Typical Uses

hydrogen-3 (tritium) 12.32 yr biochemical tracer

positron emission tomography (biomedical


carbon-11 20.33 min
imaging)

carbon-14 5.70 × 103 yr dating of artifacts

sodium-24 14.951 h cardiovascular system tracer

phosphorus-32 14.26 days biochemical tracer

potassium-40 1.248 × 109 yr dating of rocks

iron-59 44.495 days red blood cell lifetime tracer

cobalt-60 5.2712 yr radiation therapy for cancer

technetium-99m* 6.006 h biomedical imaging

iodine-131 8.0207 days thyroid studies tracer

radium-226 1.600 × 103 yr radiation therapy for cancer

uranium-238 4.468 × 109 yr dating of rocks and Earth’s crust

americium-241 432.2 yr smoke detectors

*The m denotes metastable, where an excited state nucleus decays to the ground state of the same isotope.

Radioisotope Dating Techniques


In our earlier discussion, we used the half-life of a first-order reaction to calculate how long the reaction had been occurring.
Because nuclear decay reactions follow first-order kinetics and have a rate constant that is independent of temperature and the
chemical or physical environment, we can perform similar calculations using the half-lives of isotopes to estimate the ages of
geological and archaeological artifacts. The techniques that have been developed for this application are known as radioisotope
dating techniques.
The most common method for measuring the age of ancient objects is carbon-14 dating. The carbon-14 isotope, created
continuously in the upper regions of Earth’s atmosphere, reacts with atmospheric oxygen or ozone to form 14CO2. As a result, the
CO2 that plants use as a carbon source for synthesizing organic compounds always includes a certain proportion of 14CO2
molecules as well as nonradioactive 12CO2 and 13CO2. Any animal that eats a plant ingests a mixture of organic compounds that
contains approximately the same proportions of carbon isotopes as those in the atmosphere. When the animal or plant dies, the
carbon-14 nuclei in its tissues decay to nitrogen-14 nuclei by a radioactive process known as beta decay, which releases low-energy
electrons (β particles) that can be detected and measured:
14 14 −
C → N+β (6.4.9)

The half-life for this reaction is 5700 ± 30 yr.


The 14C/12C ratio in living organisms is 1.3 × 10−12, with a decay rate of 15 dpm/g of carbon (Figure 6.4.1). Comparing the
disintegrations per minute per gram of carbon from an archaeological sample with those from a recently living sample enables
scientists to estimate the age of the artifact, illustrated below. Using this method implicitly assumes that the 14CO2/12CO2 ratio in
the atmosphere is constant, which is not strictly correct. Other methods, such as tree-ring dating, have been used to calibrate the
dates obtained by radiocarbon dating, and all radiocarbon dates reported are now corrected for minor changes in the 14CO2/12CO2
ratio over time.

6.4.2 https://chem.libretexts.org/@go/page/83322
Figure 6.4.1 : Radiocarbon Dating. A plot of the specific activity of 14C versus age for a number of archaeological samples shows
an inverse linear relationship between 14C content (a log scale) and age (a linear scale).

Example 6.4.1

In 1990, the remains of an apparently prehistoric man were found in a melting glacier in the Italian Alps. Analysis of the 14C
content of samples of wood from his tools gave a decay rate of 8.0 dpm/g carbon. How long ago did the man die?
Given: isotope and final activity
Asked for: elapsed time
Strategy:
14
A Use Equation 6.4.5 to calculate N0/N. Then substitute the value for the half-life of C into Equation 6.4.3 to find the rate
constant for the reaction.
B Using the values obtained for N0/N and the rate constant, solve Equation 6.4.8 to obtain the elapsed time.
Solution
We know the initial activity from the isotope’s identity (15 dpm/g), the final activity (8.0 dpm/g), and the half-life, so we can
use the integrated rate law for a first-order nuclear reaction (Equation 6.4.8) to calculate the elapsed time (the amount of time
elapsed since the wood for the tools was cut and began to decay).
N
ln = −kt
N0

ln(N / N0 )
=t
k

A From Equation 6.4.5, we know that A = kN. We can therefore use the initial and final activities (A0 = 15 dpm and A = 8.0
dpm) to calculate N0/N:
kN0 N0 15
df rac A0 A = = =
kN N 8.0

Now we need only calculate the rate constant for the reaction from its half-life (5730 yr) using Equation 6.4.3:
0.693
t1/2 =
k

This equation can be rearranged as follows:


0.693 0.693
−4 −1
k = = = 1.22 × 10  yr
t1/2 5730 yr

6.4.3 https://chem.libretexts.org/@go/page/83322
B Substituting into the equation for t,
ln(N0 /N ) ln(15/8.0)
3
t = = = 5.2 × 10  yr
−4
k 1.22 × 10  yr
−1

From our calculations, the man died 5200 yr ago.

Exercise 6.4.1

It is believed that humans first arrived in the Western Hemisphere during the last Ice Age, presumably by traveling over an
exposed land bridge between Siberia and Alaska. Archaeologists have estimated that this occurred about 11,000 yr ago, but
some argue that recent discoveries in several sites in North and South America suggest a much earlier arrival. Analysis of a
sample of charcoal from a fire in one such site gave a 14C decay rate of 0.4 dpm/g of carbon. What is the approximate age of
the sample?

Answer
30,000 yr

Summary
The half-life of a first-order reaction is independent of the concentration of the reactants.
The half-lives of radioactive isotopes can be used to date objects.
The half-life of a reaction is the time required for the reactant concentration to decrease to one-half its initial value. The half-life of
a first-order reaction is a constant that is related to the rate constant for the reaction: t1/2 = 0.693/k. Radioactive decay reactions are
first-order reactions. The rate of decay, or activity, of a sample of a radioactive substance is the decrease in the number of
radioactive nuclei per unit time.

This page titled 6.4: Rates of Radioactive Decay is shared under a CC BY-NC-SA 3.0 license and was authored, remixed, and/or curated by
Anonymous.

6.4.4 https://chem.libretexts.org/@go/page/83322
6.5: Stability of the Atomic Nucleus
 Learning Objectives
To calculate a mass-energy balance and a nuclear binding energy.
To understand the differences between nuclear fission and fusion.

Nuclear reactions, like chemical reactions, are accompanied by changes in energy. The energy changes in nuclear reactions,
however, are enormous compared with those of even the most energetic chemical reactions. In fact, the energy changes in a typical
nuclear reaction are so large that they result in a measurable change of mass. In this section, we describe the relationship between
mass and energy in nuclear reactions and show how the seemingly small changes in mass that accompany nuclear reactions result
in the release of enormous amounts of energy.

Mass–Energy Balance
The relationship between mass (m) and energy (E) is expressed in the following equation:
2
E = mc (6.5.1)

where c is the speed of light (2.998 × 108 m/s), and E and m are expressed in units of joules and kilograms, respectively. Albert
Einstein first derived this relationship in 1905 as part of his special theory of relativity: the mass of a particle is directly
proportional to its energy. Thus according to Equation 20.27, every mass has an associated energy, and similarly, any reaction that
involves a change in energy must be accompanied by a change in mass. This implies that all exothermic reactions should be
accompanied by a decrease in mass, and all endothermic reactions should be accompanied by an increase in mass. Given the law of
conservation of mass, how can this be true? The solution to this apparent contradiction is that chemical reactions are indeed
accompanied by changes in mass, but these changes are simply too small to be detected. As you may recall, all particles exhibit
wavelike behavior, but the wavelength is inversely proportional to the mass of the particle (actually, to its momentum, the product
of its mass and velocity). Consequently, wavelike behavior is detectable only for particles with very small masses, such as
electrons. For example, the chemical equation for the combustion of graphite to produce carbon dioxide is as follows:
1 ∘
C(graphite) + O2 (g) → C O2 (g) ΔH = −393.5 kJ/mol (6.5.2)
2

Combustion reactions are typically carried out at constant pressure, and under these conditions, the heat released or absorbed is
equal to ΔH. When a reaction is carried out at constant volume, the heat released or absorbed is equal to ΔE. For most chemical
reactions, however, ΔE ≈ ΔH. If we rewrite Einstein’s equation as
2
ΔE = (Δm)c (6.5.3)

we can rearrange the equation to obtain the following relationship between the change in mass and the change in energy:
ΔE
Δm = (6.5.4)
2
c

Because 1 J = 1 (kg·m2)/s2, the change in mass is as follows:


5 2 2
−393.5 kJ/mol −3.935 × 10 (kg ⋅ m )/(s ⋅ mol)
−12
Δm = = = −4.38 × 10  kg/mol (6.5.5)
8 2 8 2
(2.998 × 10  m/s) (2.998 × 10  m/s)

This is a mass change of about 3.6 × 10−10 g/g carbon that is burned, or about 100-millionths of the mass of an electron per atom of
carbon. In practice, this mass change is much too small to be measured experimentally and is negligible.
In contrast, for a typical nuclear reaction, such as the radioactive decay of 14C to 14N and an electron (a β particle), there is a much
larger change in mass:
14 14 0
C → N+ β (6.5.6)
−1

We can use the experimentally measured masses of subatomic particles and common isotopes to calculate the change in mass
directly. The reaction involves the conversion of a neutral 14C atom to a positively charged 14N ion (with six, not seven, electrons)
and a negatively charged β particle (an electron), so the mass of the products is identical to the mass of a neutral 14N atom. The

6.5.1 https://chem.libretexts.org/@go/page/13193
14
total change in mass during the reaction is therefore the difference between the mass of a neutral N atom (14.003074 amu) and
the mass of a 14C atom (14.003242 amu):
Δm = massproducts − massreactants (6.5.7)

= 14.003074 amu − 14.003242 amu = −0.000168 amu (6.5.8)

The difference in mass, which has been released as energy, corresponds to almost one-third of an electron. The change in mass for
the decay of 1 mol of 14C is −0.000168 g = −1.68 × 10−4 g = −1.68 × 10−7 kg. Although a mass change of this magnitude may seem
small, it is about 1000 times larger than the mass change for the combustion of graphite. The energy change is as follows:
2 −7 8 2
ΔE = (Δm)c = (−1.68 × 10  kg)(2.998 × 10  m/s) (6.5.9)
10 2 2 10 7
= −1.51 × 10 (kg ⋅ m )/ s = −1.51 × 10  J = −1.51 × 10  kJ (6.5.10)

The energy released in this nuclear reaction is more than 100,000 times greater than that of a typical chemical reaction, even
though the decay of 14C is a relatively low-energy nuclear reaction.
Because the energy changes in nuclear reactions are so large, they are often expressed in kiloelectronvolts (1 keV = 103 eV),
megaelectronvolts (1 MeV = 106 eV), and even gigaelectronvolts (1 GeV = 109 eV) per atom or particle. The change in energy that
accompanies a nuclear reaction can be calculated from the change in mass using the relationship 1 amu = 931 MeV. The energy
released by the decay of one atom of 14C is thus
931 MeV
−4
(−1.68 × 10 amu) ( ) = −0.156 MeV = −156 keV (6.5.11)
amu

 Example 6.5.1

Calculate the changes in mass (in atomic mass units) and energy (in joules per mole and electronvolts per atom) that
accompany the radioactive decay of 238U to 234Th and an α particle. The α particle absorbs two electrons from the surrounding
matter to form a helium atom.
Given: nuclear decay reaction
Asked for: changes in mass and energy
Strategy:
A Use the mass values in Table 20.1 to calculate the change in mass for the decay reaction in atomic mass units.
B Use Equation 20.30 to calculate the change in energy in joules per mole.
C Use the relationship between atomic mass units and megaelectronvolts to calculate the change in energy in electronvolts per
atom.
Solution
A Using particle and isotope masses from Table 20.1, we can calculate the change in mass as follows:
234 4 238
Δm = massproducts − massreactants = (mass Th + mass He) − mass U (6.5.12)
2

= (234.043601 amu + 4.002603 amu) − 238.050788 amu = −0.004584 amu (6.5.13)

B Thus the change in mass for 1 mol of 238U is −0.004584 g or −4.584 × 10−6 kg. The change in energy in joules per mole is as
follows:
ΔE = (Δm)c2 = (−4.584 × 10−6 kg)(2.998 × 108 m/s)2 = −4.120 × 1011 J/mol
C The change in energy in electronvolts per atom is as follows:
6
931 MeV 1 × 10  eV
−3 6
ΔE = −4.584 × 10  amu × × = −4.27 × 10  eV/atom
amu 1 MeV

6.5.2 https://chem.libretexts.org/@go/page/13193
 Exercise 6.5.1

Calculate the changes in mass (in atomic mass units) and energy (in kilojoules per mole and kiloelectronvolts per atom) that
accompany the radioactive decay of tritium (3H) to 3He and a β particle.
Answer
Δm = −2.0 × 10−5 amu; ΔE = −1.9 × 106 kJ/mol = −19 keV/atom

Nuclear Binding Energies


We have seen that energy changes in both chemical and nuclear reactions are accompanied by changes in mass. Einstein’s equation,
which allows us to interconvert mass and energy, has another interesting consequence: The mass of an atom is always less than the
sum of the masses of its component particles. The only exception to this rule is hydrogen-1 (1H), whose measured mass of
1.007825 amu is identical to the sum of the masses of a proton and an electron. In contrast, the experimentally measured mass of an
atom of deuterium (2H) is 2.014102 amu, although its calculated mass is 2.016490 amu:
m2 H = mneutron + mproton + melectron (6.5.14)

= 1.008665 amu + 1.007276 amu + 0.000549 amu = 2.016490 amu (6.5.15)

The difference between the sum of the masses of the components and the measured atomic mass is called the mass defect of the
nucleus. Just as a molecule is more stable than its isolated atoms, a nucleus is more stable (lower in energy) than its isolated
components. Consequently, when isolated nucleons assemble into a stable nucleus, energy is released. According to Equation
6.5.4, this release of energy must be accompanied by a decrease in the mass of the nucleus.

The amount of energy released when a nucleus forms from its component nucleons is the nuclear binding energy (Figure 21.2.1).
In the case of deuterium, the mass defect is 0.002388 amu, which corresponds to a nuclear binding energy of 2.22 MeV for the
deuterium nucleus. Because the magnitude of the mass defect is proportional to the nuclear binding energy, both values indicate the
stability of the nucleus.

Just as a molecule is more stable (lower in energy) than its isolated atoms, a nucleus is more stable than its isolated
components.

Figure 6.5.1 : Nuclear Binding Energy in Deuterium. The mass of a 2H atom is less than the sum of the masses of a proton, a
neutron, and an electron by 0.002388 amu; the difference in mass corresponds to the nuclear binding energy. The larger the value
of the mass defect, the greater the nuclear binding energy and the more stable the nucleus.
Not all nuclei are equally stable. Chemists describe the relative stability of different nuclei by comparing the binding energy per
nucleon, which is obtained by dividing the nuclear binding energy by the mass number (A) of the nucleus. As shown in Figure
6.5.2, the binding energy per nucleon increases rapidly with increasing atomic number until about Z = 26, where it levels off to

about 8–9 MeV per nucleon and then decreases slowly. The initial increase in binding energy is not a smooth curve but exhibits
sharp peaks corresponding to the light nuclei that have equal numbers of protons and neutrons (e.g., 4He, 12C, and 16O). As
mentioned earlier, these are particularly stable combinations.
Because the maximum binding energy per nucleon is reached at 56Fe, all other nuclei are thermodynamically unstable with regard
to the formation of 56Fe. Consequently, heavier nuclei (toward the right in 6.5.2) should spontaneously undergo reactions such as
alpha decay, which result in a decrease in atomic number. Conversely, lighter elements (on the left in Figure 6.5.2) should
spontaneously undergo reactions that result in an increase in atomic number. This is indeed the observed pattern.

6.5.3 https://chem.libretexts.org/@go/page/13193
Figure 6.5.3 : The Curve of Nuclear Binding Energy. This plot of the average binding energy per nucleon as a function of atomic
number shows that the binding energy per nucleon increases with increasing atomic number until about Z = 26, levels off, and then
decreases. The sharp peaks correspond to light nuclei that have equal numbers of protons and neutrons.

Heavier nuclei spontaneously undergo nuclear reactions that decrease their atomic number. Lighter nuclei spontaneously
undergo nuclear reactions that increase their atomic number.

 Example 6.5.2
56
Calculate the total nuclear binding energy (in megaelectronvolts) and the binding energy per nucleon for Fe. The
experimental mass of the nuclide is given in Table A4.
Given: nuclide and mass
Asked for: nuclear binding energy and binding energy per nucleon
Strategy:
A Sum the masses of the protons, electrons, and neutrons or, alternatively, use the mass of the appropriate number of 1H atoms
(because its mass is the same as the mass of one electron and one proton).
B Calculate the mass defect by subtracting the experimental mass from the calculated mass.
C Determine the nuclear binding energy by multiplying the mass defect by the change in energy in electronvolts per atom.
Divide this value by the number of nucleons to obtain the binding energy per nucleon.
Solution
A An iron-56 atom has 26 protons, 26 electrons, and 30 neutrons. We could add the masses of these three sets of particles;
however, noting that 26 protons and 26 electrons are equivalent to 26 1H atoms, we can calculate the sum of the masses more
quickly as follows:
1 1
calculated mass = 26(mass  H) + 30(mass  n) (6.5.16)
1 0

= 26(1.007825)amu + 30(1.008665)amu = 56.463400 amu (6.5.17)

experimental mass = 55.934938 (6.5.18)

B We subtract to find the mass defect:


mass defect = calculated mass − experimental mass (6.5.19)

= 56.463400 amu − 55.934938 amu = 0.528462 amu (6.5.20)

C The nuclear binding energy is thus 0.528462 amu × 931 MeV/amu = 492 MeV. The binding energy per nucleon is 492
MeV/56 nucleons = 8.79 MeV/nucleon.

6.5.4 https://chem.libretexts.org/@go/page/13193
 Exercise 6.5.2

Calculate the total nuclear binding energy (in megaelectronvolts) and the binding energy per nucleon for 238U.
Answer 1800 MeV/238U; 7.57 MeV/nucleon

Nuclear Fission and Fusion


Nuclear fission is the splitting of a heavy nucleus into two lighter ones. Fission was discovered in 1938 by the German scientists
Otto Hahn, Lise Meitner, and Fritz Strassmann, who bombarded a sample of uranium with neutrons in an attempt to produce new
elements with Z > 92. They observed that lighter elements such as barium (Z = 56) were formed during the reaction, and they
realized that such products had to originate from the neutron-induced fission of uranium-235:
235 1 141 92 1
U+ n→ Ba + Kr + 30 n (6.5.21)
92 0 56 36

This hypothesis was confirmed by detecting the krypton-92 fission product. The nucleus usually divides asymmetrically rather than
into two equal parts, and the fission of a given nuclide does not give the same products every time.
In a typical nuclear fission reaction, more than one neutron is released by each dividing nucleus. When these neutrons collide with
and induce fission in other neighboring nuclei, a self-sustaining series of nuclear fission reactions known as a nuclear chain
reaction can result (Figure 6.5.3). For example, the fission of 235U releases two to three neutrons per fission event. If absorbed by
other 235U nuclei, those neutrons induce additional fission events, and the rate of the fission reaction increases geometrically. Each
series of events is called a generation. Experimentally, it is found that some minimum mass of a fissile isotope is required to sustain
a nuclear chain reaction; if the mass is too low, too many neutrons are able to escape without being captured and inducing a fission
reaction. The minimum mass capable of supporting sustained fission is called the critical mass. This amount depends on the purity
of the material and the shape of the mass, which corresponds to the amount of surface area available from which neutrons can
escape, and on the identity of the isotope. If the mass of the fissile isotope is greater than the critical mass, then under the right
conditions, the resulting supercritical mass can release energy explosively. The enormous energy released from nuclear chain
reactions is responsible for the massive destruction caused by the detonation of nuclear weapons such as fission bombs, but it also
forms the basis of the nuclear power industry.

Figure 6.5.3 : A Nuclear Chain Reaction. The process is initiated by the collision of a single neutron with a 235U nucleus, which
undergoes fission, as shown in Figure 20.6. Because each neutron released can cause the fission of another 235U nucleus, the rate of
a fission reaction accelerates geometrically. Each series of events is a generation.
Nuclear fusion, in which two light nuclei combine to produce a heavier, more stable nucleus, is the opposite of nuclear fission. As
in the nuclear transmutation reactions discussed in Section 20.2, the positive charge on both nuclei results in a large electrostatic
energy barrier to fusion. This barrier can be overcome if one or both particles have sufficient kinetic energy to overcome the
electrostatic repulsions, allowing the two nuclei to approach close enough for a fusion reaction to occur. The principle is similar to
adding heat to increase the rate of a chemical reaction. As shown in the plot of nuclear binding energy per nucleon versus atomic
number in Figure 21.2.3, fusion reactions are most exothermic for the lightest element. For example, in a typical fusion reaction,
two deuterium atoms combine to produce helium-3, a process known as deuterium–deuterium fusion (D–D fusion):
2 3 1
21 H → He + n (6.5.22)
2 0

6.5.5 https://chem.libretexts.org/@go/page/13193
In another reaction, a deuterium atom and a tritium atom fuse to produce helium-4 (Figure 6.5.4), a process known as deuterium–
tritium fusion (D–T fusion):
2 3 4 1
H+ H → He + n (6.5.23)
1 1 2 0

Figure 6.5.4 : Nuclear Fusion. In a nuclear fusion reaction, lighter nuclei combine to produce a heavier nucleus. As shown, fusion
of 3H and 2H to give 4He and a neutron releases an enormous amount of energy. In principle, nuclear fusion can produce much
more energy than fission, but very high kinetic energy is required to overcome electrostatic repulsions between the positively
charged nuclei and initiate the fusion reaction.
Initiating these reactions, however, requires a temperature comparable to that in the interior of the sun (approximately 1.5 × 107 K).
Currently, the only method available on Earth to achieve such a temperature is the detonation of a fission bomb. For example, the
so-called hydrogen bomb (or H bomb) is actually a deuterium–tritium bomb (a D–T bomb), which uses a nuclear fission reaction to
create the very high temperatures needed to initiate fusion of solid lithium deuteride (6LiD), which releases neutrons that then react
with 6Li, producing tritium. The deuterium-tritium reaction releases energy explosively. Example 9 and its corresponding exercise
demonstrate the enormous amounts of energy produced by nuclear fission and fusion reactions. In fact, fusion reactions are the
power sources for all stars, including our sun.

 Example 6.5.3

Calculate the amount of energy (in electronvolts per atom and kilojoules per mole) released when the neutron-induced fission
of 235U produces 144Cs, 90Rb, and two neutrons:
235 1 144 90 1
U+ n→ Cs + Rb + 2 n
92 0 55 37 0

Given: balanced nuclear reaction


Asked for: energy released in electronvolts per atom and kilojoules per mole
Strategy:
A Following the method used in Example 6.5.1, calculate the change in mass that accompanies the reaction. Convert this value
to the change in energy in electronvolts per atom.
B Calculate the change in mass per mole of 235U. Then use Equation 6.5.3 to calculate the change in energy in kilojoules per
mole.
Solution
A The change in mass that accompanies the reaction is as follows:
144 90 1 235
Δm = massproducts − massreactants = mass( Cs + Rb + n) − mass  U (6.5.24)
55 37 0 92

= (143.932077 amu + 89.914802 amu + 1.008665 amu) − 235.043930 amu (6.5.25)

= −0.188386 amu (6.5.26)

The change in energy in electronvolts per atom is as follows:


ΔE = (−0.188386 amu)(931 MeV/amu) = −175 MeV

B The change in mass per mole of 235


92
U is −0.188386 g = −1.88386 × 10−4 kg, so the change in energy in kilojoules per mole is
as follows:
2 −4 8 2
ΔE = (Δm)c = (−1.88386 × 10  kg)(2.998 × 10  m/s) (6.5.27)

13 10
= −1.693 × 10  J/mol = −1.693 × 10  kJ/mol (6.5.28)

6.5.6 https://chem.libretexts.org/@go/page/13193
 Exercise 6.5.3
Calculate the amount of energy (in electronvolts per atom and kilojoules per mole) released when deuterium and tritium fuse to
give helium-4 and a neutron:
2 3 4 1
H+ H → He + n
1 1 2 0

Answer
ΔE = −17.6 MeV/atom = −1.697 × 109 kJ/mol

Summary
Unlike a chemical reaction, a nuclear reaction results in a significant change in mass and an associated change of energy, as
described by Einstein’s equation. Nuclear reactions are accompanied by large changes in energy, which result in detectable changes
in mass. The change in mass is related to the change in energy according to Einstein’s equation: ΔE = (Δm)c2. Large changes in
energy are usually reported in kiloelectronvolts or megaelectronvolts (thousands or millions of electronvolts). With the exception of
1H, the experimentally determined mass of an atom is always less than the sum of the masses of the component particles (protons,

neutrons, and electrons) by an amount called the mass defect of the nucleus. The energy corresponding to the mass defect is the
nuclear binding energy, the amount of energy released when a nucleus forms from its component particles. In nuclear fission,
nuclei split into lighter nuclei with an accompanying release of multiple neutrons and large amounts of energy. The critical mass is
the minimum mass required to support a self-sustaining nuclear chain reaction. Nuclear fusion is a process in which two light
nuclei combine to produce a heavier nucleus plus a great deal of energy.

6.5: Stability of the Atomic Nucleus is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
24.5: Thermodynamic Stability of the Atomic Nucleus by Anonymous is licensed CC BY-NC-SA 4.0.

6.5.7 https://chem.libretexts.org/@go/page/13193
6.6: The Origin of the Elements
 Learning Objectives
To understand how nuclear transmutation reactions lead to the formation of the elements in stars and how they can be used
to synthesize transuranium elements.

The relative abundances of the elements in the known universe vary by more than 12 orders of magnitude. For the most part, these
differences in abundance cannot be explained by differences in nuclear stability. Although the 56Fe nucleus is the most stable
nucleus known, the most abundant element in the known universe is not iron, but hydrogen (1H), which accounts for about 90% of
all atoms. In fact, 1H is the raw material from which all other elements are formed. In this section, we explain why 1H and 2He
together account for at least 99% of all the atoms in the known universe. We also describe the nuclear reactions that take place in
stars, which transform one nucleus into another and create all the naturally occurring elements.

Relative Abundances of the Elements on Earth and in the Known Universe


The relative abundances of the elements in the known universe and on Earth relative to silicon are shown in Figure 6.6.1. The data
are estimates based on the characteristic emission spectra of the elements in stars, the absorption spectra of matter in clouds of
interstellar dust, and the approximate composition of Earth as measured by geologists. The data in Figure 6.6.1 illustrate two
important points. First, except for hydrogen, the most abundant elements have even atomic numbers. Not only is this consistent
with the known trends in nuclear stability, but it also suggests that heavier elements are formed by combining helium nuclei (Z =
2). Second, the relative abundances of the elements in the known universe and on Earth are often very different, as indicated by the
data in Table 6.6.1 for some common elements.

Figure 6.6.1 : The Relative Abundances of the Elements in the Universe and on Earth. In this logarithmic plot, the relative
abundances of the elements relative to that of silicon (arbitrarily set equal to 1) in the universe (green bars) and on Earth (purple
bars) are shown as a function of atomic number. Elements with even atomic numbers are generally more abundant in the universe
than elements with odd atomic numbers. Also, the relative abundances of many elements in the universe are very different from
their relative abundances on Earth. (CC BY-NC-SA 3.0; anonymous)
Bar graph of abundance realtive to silicon in pbb against atomic number. The purple bars are universe elements while the gray bars
are terrestrial elements.
Some of these differences are easily explained. For example, nonmetals such as H, He, C, N, O, Ne, and Kr are much less abundant
relative to silicon on Earth than they are in the rest of the universe. These elements are either noble gases (He, Ne, and Kr) or
elements that form volatile hydrides, such as NH3, CH4, and H2O. Because Earth’s gravity is not strong enough to hold such light
substances in the atmosphere, these elements have been slowly diffusing into outer space ever since our planet was formed. Argon
is an exception; it is relatively abundant on Earth compared with the other noble gases because it is continuously produced in rocks
by the radioactive decay of isotopes such as 40K. In contrast, many metals, such as Al, Na, Fe, Ca, Mg, K, and Ti, are relatively
abundant on Earth because they form nonvolatile compounds, such as oxides, that cannot escape into space. Other metals, however,
are much less abundant on Earth than in the universe; some examples are Ru and Ir. This section explains some of the reasons for
the great differences in abundances of the metallic elements.

6.6.1 https://chem.libretexts.org/@go/page/13195
Table 6.6.1 : Relative Abundances of Elements on Earth and in the Known Universe
Terrestrial/Universal Element Abundance Ratio

H 0.0020

He 2.4 × 10−8

C 0.36

N 0.02

O 46

Ne 1.9 × 10−6

Na 1200

Mg 48

Al 1600

Si 390

S 0.84

K 5000

Ca 710

Ti 2200

Fe 57

All the elements originally present on Earth (and on other planets) were synthesized from hydrogen and helium nuclei in the
interiors of stars that have long since exploded and disappeared. Six of the most abundant elements in the universe (C, O, Ne, Mg,
Si, and Fe) have nuclei that are integral multiples of the helium-4 nucleus, which suggests that helium-4 is the primary building
block for heavier nuclei.

Synthesis of the Elements in Stars


Elements are synthesized in discrete stages during the lifetime of a star, and some steps occur only in the most massive stars known
(Figure 6.6.2). Initially, all stars are formed by the aggregation of interstellar “dust,” which is mostly hydrogen. As the cloud of
dust slowly contracts due to gravitational attraction, its density eventually reaches about 100 g/cm3, and the temperature increases
to about 1.5 × 107 K, forming a dense plasma of ionized hydrogen nuclei. At this point, self-sustaining nuclear reactions begin, and
the star “ignites,” creating a yellow star like our sun.

6.6.2 https://chem.libretexts.org/@go/page/13195
Figure 6.6.2 : Nuclear Reactions during the Life Cycle of a Massive Star. At each stage in the lifetime of a star, a different fuel is
used for nuclear fusion, resulting in the formation of different elements. Fusion of hydrogen to give helium is the primary fusion
reaction in young stars. As the star ages, helium accumulates and begins to “burn,” undergoing fusion to form heavier elements
such as carbon and oxygen. As the adolescent star matures, significant amounts of iron and nickel are formed by fusion of the
heavier elements formed previously. The heaviest elements are formed only during the final death throes of the star—the formation
of a nova or supernova.
The stages in a stare lifetime are yellow star, red giant, red supergiant, massive red supergiant and finally supernova.
In the first stage of its life, the star is powered by a series of nuclear fusion reactions that convert hydrogen to helium:
1 1 2 0
H+ H → H+ β (6.6.1)
1 1 1 +1

2 1 3 0
H+ H → He + γ
1 1 2 0

3 3 4 1
He + He → He + 21 H
2 2 2

The overall reaction is the conversion of four hydrogen nuclei to a helium-4 nucleus, which is accompanied by the release of two
positrons, two γ rays, and a great deal of energy:
1 4 0 0
41 H → He + 2+1 β + 20 γ (6.6.2)
2

These reactions are responsible for most of the enormous amount of energy that is released as sunlight and solar heat. It takes
several billion years, depending on the size of the star, to convert about 10% of the hydrogen to helium.
Once large amounts of helium-4 have been formed, they become concentrated in the core of the star, which slowly becomes denser
and hotter. At a temperature of about 2 × 108 K, the helium-4 nuclei begin to fuse, producing beryllium-8:
4 8
2 He → Be (6.6.3)
2 4

Although beryllium-8 has both an even mass number and an even atomic number, its also has a low neutron-to-proton ratio (and
other factors beyond the scope of this text) that makes it unstable; it decomposes in only about 10−16 s. Nonetheless, this is long
enough for it to react with a third helium-4 nucleus to form carbon-12, which is very stable. Sequential reactions of carbon-12 with
helium-4 produce the elements with even numbers of protons and neutrons up to magnesium-24:
4 4 4 4
He He He He
2 2 2 2
8 12 16 20 24
Be −→
− C −→
− O −→
− Ne −→
− Mg (6.6.4)
4 6 8 10 12

So much energy is released by these reactions that it causes the surrounding mass of hydrogen to expand, producing a red giant that
is about 100 times larger than the original yellow star.
As the star expands, heavier nuclei accumulate in its core, which contracts further to a density of about 50,000 g/cm3, so the core
becomes even hotter. At a temperature of about 7 × 108 K, carbon and oxygen nuclei undergo nuclear fusion reactions to produce
sodium and silicon nuclei:
12 12 23 1
C+ C → Na + H (6.6.5)
6 6 11 1

12 16 28 0
C+ O→ Si + γ (6.6.6)
6 8 14 0

6.6.3 https://chem.libretexts.org/@go/page/13195
At these temperatures, carbon-12 reacts with helium-4 to initiate a series of reactions that produce more oxygen-16, neon-20,
magnesium-24, and silicon-28, as well as heavier nuclides such as sulfur-32, argon-36, and calcium-40:
4 4 4 4 4 4 4
He He He He He He He
2 2 2 2 2 2 2
12 16 20 24 28 32 36 40
C −→
− O −→
− Ne −→
− Mg −→
− Si −→
− S −→
− Ar −→
− Ca (6.6.7)
6 8 10 12 14 16 18 20

The energy released by these reactions causes a further expansion of the star to form a red supergiant, and the core temperature
increases steadily. At a temperature of about 3 × 109 K, the nuclei that have been formed exchange protons and neutrons freely.
This equilibration process forms heavier elements up to iron-56 and nickel-58, which have the most stable nuclei known.

The Formation of Heavier Elements in Supernovas


None of the processes described so far produces nuclei with Z > 28. All naturally occurring elements heavier than nickel are
formed in the rare but spectacular cataclysmic explosions called supernovas (Figure 6.6.2). When the fuel in the core of a very
massive star has been consumed, its gravity causes it to collapse in about 1 s. As the core is compressed, the iron and nickel nuclei
within it disintegrate to protons and neutrons, and many of the protons capture electrons to form neutrons. The resulting neutron
star is a dark object that is so dense that atoms no longer exist. Simultaneously, the energy released by the collapse of the core
causes the supernova to explode in what is arguably the single most violent event in the universe. The force of the explosion blows
most of the star’s matter into space, creating a gigantic and rapidly expanding dust cloud, or nebula (Figure 6.6.3). During the
extraordinarily short duration of this event, the concentration of neutrons is so great that multiple neutron-capture events occur,
leading to the production of the heaviest elements and many of the less stable nuclides. Under these conditions, for example, an
iron-56 nucleus can absorb as many as 64 neutrons, briefly forming an extraordinarily unstable iron isotope that can then undergo
multiple rapid β-decay processes to produce tin-120:
56 1 120 120 0
Fe + 640 n → Fe → Sn + 24−1 β (6.6.8)
26 26 50

Figure 6.6.3 : A Supernova. A view of the remains of Supernova 1987A, located in the Large Magellanic Cloud, showing the
circular halo of expanding debris produced by the explosion. Multiple neutron-capture events occur during a supernova explosion,
forming both the heaviest elements and many of the less stable nuclides.
Although a supernova occurs only every few hundred years in a galaxy such as the Milky Way, these rare explosions provide the
only conditions under which elements heavier than nickel can be formed. The force of the explosions distributes these elements
throughout the galaxy surrounding the supernova, and eventually they are captured in the dust that condenses to form new stars.
Based on its elemental composition, our sun is thought to be a second- or third-generation star. It contains a considerable amount of
cosmic debris from the explosion of supernovas in the remote past.

 Example 6.6.1: Carbon Burning Stars


The reaction of two carbon-12 nuclei in a carbon-burning star can produce elements other than sodium. Write a balanced
nuclear equation for the formation of
a. magnesium-24.
b. neon-20 from two carbon-12 nuclei.
Given: reactant and product nuclides
Asked for: balanced nuclear equation

6.6.4 https://chem.libretexts.org/@go/page/13195
Strategy:
Use conservation of mass and charge to determine the type of nuclear reaction that will convert the reactant to the indicated
product. Write the balanced nuclear equation for the reaction.

Solution
a. A magnesium-24 nucleus (Z = 12, A = 24) has the same nucleons as two carbon-12 nuclei (Z = 6, A = 12). The reaction is
therefore a fusion of two carbon-12 nuclei, and no other particles are produced: C + C → Mg .
12
6
12
6
24
12

b. The neon-20 product has Z = 10 and A = 20. The conservation of mass requires that the other product have A = (2 × 12) −
20 = 4; because of conservation of charge, it must have Z = (2 × 6) − 10 = 2. These are the characteristics of an α particle.
The reaction is therefore C + C → Ne + α .
12
6
12
6
20
10
4
2

 Exercise 6.6.1

How many neutrons must an iron-56 nucleus absorb during a supernova explosion to produce an arsenic-75 nucleus? Write a
balanced nuclear equation for the reaction.

Answer
19 neutrons; 56
26
1
Fe + 190 n →
75
26
Fe →
75
33
0
As + 7−1 β

Summary
Hydrogen and helium are the most abundant elements in the universe. Heavier elements are formed in the interior of stars via
multiple neutron-capture events. By far the most abundant element in the universe is hydrogen. The fusion of hydrogen nuclei to
form helium nuclei is the major process that fuels young stars such as the sun. Elements heavier than helium are formed from
hydrogen and helium in the interiors of stars. Successive fusion reactions of helium nuclei at higher temperatures create elements
with even numbers of protons and neutrons up to magnesium and then up to calcium. Eventually, the elements up to iron-56 and
nickel-58 are formed by exchange processes at even higher temperatures. Heavier elements can only be made by a process that
involves multiple neutron-capture events, which can occur only during the explosion of a supernova.

6.6: The Origin of the Elements is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
21.3: Nuclear Transmutations by Anonymous is licensed CC BY-NC-SA 3.0.

6.6.5 https://chem.libretexts.org/@go/page/13195
6.7: Transmutation of the Elements
Although the conversion of one element to another is the basis of natural radioactive decay, it is also possible to convert one
element to another experimentally. The conversion of one element to another is the process of transmutation. In 1919 Rutherford
converted a stable isotope to an unstable one. By bombarding N with α particles he created O, which is an unstable isotope of
14 17

oxygen. Transmutation may also be accomplished by bombardment with neutrons.


14 4 17 1
N + 2 He → 8
O + 1H (6.7.1)
7

Historically, part of Alchemy was the study of methods of creating gold from base metals, such lead. Where the Alchemists failed
in this quest, we can now succeed. Thus, bombardment of platinum-198 with a neutron creates an unstable isotope of platinum that
undergoes beta decay to gold-199. Unfortunately, while we may succeed in making gold, the platinum we make it from is actually
worth more than the gold making this particular transmutation economically non-viable!
198 1 199 199 0
Pt + n → Pt → 79
Au + −1 β (6.7.2)
78 0 78

Transuranium Elements
Uranium (Z = 92) is the heaviest naturally occurring element. Consequently, all the elements with Z > 92, the transuranium
elements, are artificial and have been prepared by bombarding suitable target nuclei with smaller particles. The first of the
transuranium elements to be prepared was neptunium (Z = 93), which was synthesized in 1940 by bombarding a 238U target with
neutrons. As shown in Equation \(\ref{20.41\) this reaction occurs in two steps. Initially, a neutron combines with a 238U nucleus to
form 239U, which is unstable and undergoes beta decay to produce 239Np:
238 1 239 239 0
U+ n→ U → Np + β (6.7.3)
92 0 92 93 −1

Subsequent beta decay of 239Np produces the second transuranium element, plutonium (Z = 94):
239 239 0
Np → Pu + β (6.7.4)
93 94 −1

Bombarding the target with more massive nuclei creates elements that have atomic numbers significantly greater than that of the
target nucleus (Table 6.7.1). Such techniques have resulted in the creation of the superheavy elements 114 and 116, both of which
lie in or near the “island of stability." As of this writing, 22 transuranium elements have been produced and officially recognized by
IUPAC; several other elements have formation claims that are waiting for approval.
Table 6.7.1 : Preparation of Some of the Transuranium Elements
Atomic
Name Symbol Reaction
Number

americiu
Am 95 239
Pu +
1
n ⟶
240
Am +
0
e
m 94 0 95 −1

curium Cm 96 239
94
Pu +
4
2
He ⟶
242
96
Cm +
1
0
n

californiu
Cf 98 242
Cm +
4
He ⟶
243
Bk + 2
1
n
m 96 2 97 0

einsteiniu
Es 99 238
U + 15
1
n ⟶
253
Es + 7
0
e
m 92 0 99 −1

mendelev
Md 101 253
Es +
4
He ⟶
256
Md +
1
n
ium 99 2 101 0

nobelium No 102 246


96
Cm +
12
6
C ⟶
254
102
No + 4
1
0
n

rutherfor
Rf 104 249
Cf +
12
C ⟶
257
Rf + 4
1
n
dium 98 6 104 0

206 54 257
seaborgiu Pb + Cr ⟶ Sg + 3
1
n
Sg 106 82 24 106 0

m 249
98
Cf +
18
8
O ⟶
263
106
Sg + 4
1
0
n

meitneriu
Mt 107 209
Bi +
58
Fe ⟶
266
Mt +
1
n
m 83 26 109 0

6.7.1 https://chem.libretexts.org/@go/page/13196
Particle Accelerators
A device called a particle accelerator is used to accelerate positively charged particles to the speeds needed to overcome the
electrostatic repulsions between them and the target nuclei by using electrical and magnetic fields. Operationally, the simplest
particle accelerator is the linear accelerator (Figure 6.7.2), in which a beam of particles is injected at one end of a long evacuated
tube. Rapid alternation of the polarity of the electrodes along the tube causes the particles to be alternately accelerated toward a
region of opposite charge and repelled by a region with the same charge, resulting in a tremendous acceleration as the particle
travels down the tube. A modern linear accelerator such as the Stanford Linear Accelerator (SLAC) at Stanford University is about
2 miles long.

Figure 6.7.2 : A Linear Particle Accelerator. (a) An aerial view of the SLAC, the longest linear particle accelerator in the world; the
overall length of the tunnel is 2 miles. (b) Rapidly reversing the polarity of the electrodes in the tube causes the charged particles to
be alternately attracted as they enter one section of the tube and repelled as they leave that section. As a result, the particles are
continuously accelerated along the length of the tube.
To achieve the same outcome in less space, a particle accelerator called a cyclotron forces the charged particles to travel in a
circular path rather than a linear one. The particles are injected into the center of a ring and accelerated by rapidly alternating the
polarity of two large D-shaped electrodes above and below the ring, which accelerates the particles outward along a spiral path
toward the target.
The length of a linear accelerator and the size of the D-shaped electrodes in a cyclotron severely limit the kinetic energy that
particles can attain in these devices. These limitations can be overcome by using a synchrotron, a hybrid of the two designs. A
synchrotron contains an evacuated tube similar to that of a linear accelerator, but the tube is circular and can be more than a mile in
diameter. Charged particles are accelerated around the circle by a series of magnets whose polarities rapidly alternate.

Contributors and Attributions


Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Hans Lohninger (Epina eBook Team)
Andrew R. Barron

6.7: Transmutation of the Elements is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

6.7.2 https://chem.libretexts.org/@go/page/13196
6.8: Nuclear Fission
Skills to Develop
Explain nuclear fission
Relate the concepts of critical mass and nuclear chain reactions
Summarize basic requirements for nuclear fission

Many heavier elements with smaller binding energies per nucleon can decompose into more stable elements that have intermediate mass numbers
and larger binding energies per nucleon—that is, mass numbers and binding energies per nucleon that are closer to the “peak” of the binding energy
graph near 56. Sometimes neutrons are also produced. This decomposition is called fission, the breaking of a large nucleus into smaller pieces. The
breaking is rather random with the formation of a large number of different products. Fission usually does not occur naturally, but is induced by
bombardment with neutrons. The first reported nuclear fission occurred in 1939 when three German scientists, Lise Meitner, Otto Hahn, and Fritz
Strassman, bombarded uranium-235 atoms with slow-moving neutrons that split the U-238 nuclei into smaller fragments that consisted of several
neutrons and elements near the middle of the periodic table. Since then, fission has been observed in many other isotopes, including most actinide
isotopes that have an odd number of neutrons. A typical nuclear fission reaction is shown in Figure 6.8.1.

Figure 6.8.1 : When a slow neutron hits a fissionable U-235 nucleus, it is absorbed and forms an unstable U-236 nucleus. The U-236 nucleus then
rapidly breaks apart into two smaller nuclei (in this case, Ba-141 and Kr-92) along with several neutrons (usually two or three), and releases a very
large amount of energy.
Among the products of Meitner, Hahn, and Strassman’s fission reaction were barium, krypton, lanthanum, and cerium, all of which have nuclei that
are more stable than uranium-235. Since then, hundreds of different isotopes have been observed among the products of fissionable substances. A
few of the many reactions that occur for U-235, and a graph showing the distribution of its fission products and their yields, are shown in Figure
6.8.2.. Similar fission reactions have been observed with other uranium isotopes, as well as with a variety of other isotopes such as those of

plutonium.

Figure 6.8.2 : (a) Nuclear fission of U-235 produces a range of fission products. (b) The larger fission products of U-235 are typically one isotope
with a mass number around 85–105, and another isotope with a mass number that is about 50% larger, that is, about 130–150.
A tremendous amount of energy is produced by the fission of heavy elements. For instance, when one mole of U-235 undergoes fission, the products
weigh about 0.2 grams less than the reactants; this “lost” mass is converted into a very large amount of energy, about 1.8 × 1010 kJ per mole of U-
235. Nuclear fission reactions produce incredibly large amounts of energy compared to chemical reactions. The fission of 1 kilogram of uranium-
235, for example, produces about 2.5 million times as much energy as is produced by burning 1 kilogram of coal.
As described earlier, when undergoing fission U-235 produces two “medium-sized” nuclei, and two or three neutrons. These neutrons may then
cause the fission of other uranium-235 atoms, which in turn provide more neutrons that can cause fission of even more nuclei, and so on. If this
occurs, we have a nuclear chain reaction (Figure 6.8.3). On the other hand, if too many neutrons escape the bulk material without interacting with a
nucleus, then no chain reaction will occur.

6.8.1 https://chem.libretexts.org/@go/page/83458
Figure 6.8.3 : The fission of a large nucleus, such as U-235, produces two or three neutrons, each of which is capable of causing fission of another
nucleus by the reactions shown. If this process continues, a nuclear chain reaction occurs.
Material that can sustain a nuclear fission chain reaction is said to be fissile or fissionable. (Technically, fissile material can undergo fission with
neutrons of any energy, whereas fissionable material requires high-energy neutrons.) Nuclear fission becomes self-sustaining when the number of
neutrons produced by fission equals or exceeds the number of neutrons absorbed by splitting nuclei plus the number that escape into the
surroundings. The amount of a fissionable material that will support a self-sustaining chain reaction is a critical mass. An amount of fissionable
material that cannot sustain a chain reaction is a subcritical mass. An amount of material in which there is an increasing rate of fission is known as a
supercritical mass. The critical mass depends on the type of material: its purity, the temperature, the shape of the sample, and how the neutron
reactions are controlled (Figure 6.8.4).

Figure 6.8.4 : (a) In a subcritical mass, the fissile material is too small and allows too many neutrons to escape the material, so a chain reaction does
not occur. (b) In a critical mass, a large enough number of neutrons in the fissile material induce fission to create a chain reaction.
An atomic bomb (Figure 6.8.5) contains several pounds of fissionable material, U or Pu, a source of neutrons, and an explosive device for
235
92
239
94

compressing it quickly into a small volume. When fissionable material is in small pieces, the proportion of neutrons that escape through the
relatively large surface area is great, and a chain reaction does not take place. When the small pieces of fissionable material are brought together
quickly to form a body with a mass larger than the critical mass, the relative number of escaping neutrons decreases, and a chain reaction and
explosion result.

6.8.2 https://chem.libretexts.org/@go/page/83458
Figure 6.8.5 : (a) The nuclear fission bomb that destroyed Hiroshima on August 6, 1945, consisted of two subcritical masses of U-235, where
conventional explosives were used to fire one of the subcritical masses into the other, creating the critical mass for the nuclear explosion. (b) The
plutonium bomb that destroyed Nagasaki on August 12, 1945, consisted of a hollow sphere of plutonium that was rapidly compressed by
conventional explosives. This led to a concentration of plutonium in the center that was greater than the critical mass necessary for the nuclear
explosion.

Fission Reactors
Chain reactions of fissionable materials can be controlled and sustained without an explosion in a nuclear reactor (Figure 6.8.6). Any nuclear reactor
that produces power via the fission of uranium or plutonium by bombardment with neutrons must have at least five components: nuclear fuel
consisting of fissionable material, a nuclear moderator, reactor coolant, control rods, and a shield and containment system. We will discuss these
components in greater detail later in the section. The reactor works by separating the fissionable nuclear material such that a critical mass cannot be
formed, controlling both the flux and absorption of neutrons to allow shutting down the fission reactions. In a nuclear reactor used for the production
of electricity, the energy released by fission reactions is trapped as thermal energy and used to boil water and produce steam. The steam is used to
turn a turbine, which powers a generator for the production of electricity.

Figure 6.8.6 : (a) The Diablo Canyon Nuclear Power Plant near San Luis Obispo is the only nuclear power plant currently in operation in California.
The domes are the containment structures for the nuclear reactors, and the brown building houses the turbine where electricity is generated. Ocean
water is used for cooling. (b) The Diablo Canyon uses a pressurized water reactor, one of a few different fission reactor designs in use around the
world, to produce electricity. Energy from the nuclear fission reactions in the core heats water in a closed, pressurized system. Heat from this system
produces steam that drives a turbine, which in turn produces electricity. (credit a: modification of work by “Mike” Michael L. Baird; credit b:
modification of work by the Nuclear Regulatory Commission)

Nuclear Fuels
Nuclear fuel consists of a fissionable isotope, such as uranium-235, which must be present in sufficient quantity to provide a self-sustaining chain
reaction. In the United States, uranium ores contain from 0.05–0.3% of the uranium oxide U3O8; the uranium in the ore is about 99.3%
nonfissionable U-238 with only 0.7% fissionable U-235. Nuclear reactors require a fuel with a higher concentration of U-235 than is found in nature;
it is normally enriched to have about 5% of uranium mass as U-235. At this concentration, it is not possible to achieve the supercritical mass
necessary for a nuclear explosion. Uranium can be enriched by gaseous diffusion (the only method currently used in the US), using a gas centrifuge,
or by laser separation.
In the gaseous diffusion enrichment plant where U-235 fuel is prepared, UF6 (uranium hexafluoride) gas at low pressure moves through barriers that
have holes just barely large enough for UF6 to pass through. The slightly lighter 235UF6 molecules diffuse through the barrier slightly faster than the
heavier 238UF6 molecules. This process is repeated through hundreds of barriers, gradually increasing the concentration of 235UF6 to the level needed
by the nuclear reactor. The basis for this process, Graham’s law, is described in the chapter on gases. The enriched UF6 gas is collected, cooled until
it solidifies, and then taken to a fabrication facility where it is made into fuel assemblies. Each fuel assembly consists of fuel rods that contain many
thimble-sized, ceramic-encased, enriched uranium (usually UO2) fuel pellets. Modern nuclear reactors may contain as many as 10 million fuel
pellets. The amount of energy in each of these pellets is equal to that in almost a ton of coal or 150 gallons of oil.

6.8.3 https://chem.libretexts.org/@go/page/83458
Nuclear Moderators
Neutrons produced by nuclear reactions move too fast to cause fission (Figure 6.8.4). They must first be slowed to be absorbed by the fuel and
produce additional nuclear reactions. A nuclear moderator is a substance that slows the neutrons to a speed that is low enough to cause fission. Early
2
reactors used high-purity graphite as a moderator. Modern reactors in the US exclusively use heavy water ( H O) or light water (ordinary H2O),
1 2

whereas some reactors in other countries use other materials, such as carbon dioxide, beryllium, or graphite.

Reactor Coolants
A nuclear reactor coolant is used to carry the heat produced by the fission reaction to an external boiler and turbine, where it is transformed into
electricity. Two overlapping coolant loops are often used; this counteracts the transfer of radioactivity from the reactor to the primary coolant loop.
All nuclear power plants in the US use water as a coolant. Other coolants include molten sodium, lead, a lead-bismuth mixture, or molten salts.

Control Rods
Nuclear reactors use control rods (Figure 6.8.8) to control the fission rate of the nuclear fuel by adjusting the number of slow neutrons present to
keep the rate of the chain reaction at a safe level. Control rods are made of boron, cadmium, hafnium, or other elements that are able to absorb
neutrons. Boron-10, for example, absorbs neutrons by a reaction that produces lithium-7 and alpha particles:
10 1 7 4
5
B+ n ⟶ 3
Li + 2 He (6.8.1)
0

When control rod assemblies are inserted into the fuel element in the reactor core, they absorb a larger fraction of the slow neutrons, thereby slowing
the rate of the fission reaction and decreasing the power produced. Conversely, if the control rods are removed, fewer neutrons are absorbed, and the
fission rate and energy production increase. In an emergency, the chain reaction can be shut down by fully inserting all of the control rods into the
nuclear core between the fuel rods.

Figure 6.8.7 : The nuclear reactor core shown in (a) contains the fuel and control rod assembly shown in (b). (credit: modification of work by E.
Generalic)

Shield and Containment System


During its operation, a nuclear reactor produces neutrons and other radiation. Even when shut down, the decay products are radioactive. In addition,
an operating reactor is thermally very hot, and high pressures result from the circulation of water or another coolant through it. Thus, a reactor must
withstand high temperatures and pressures, and must protect operating personnel from the radiation. Reactors are equipped with a containment
system (or shield) that consists of three parts:
1. The reactor vessel, a steel shell that is 3–20-centimeters thick and, with the moderator, absorbs much of the radiation produced by the reactor
2. A main shield of 1–3 meters of high-density concrete
3. A personnel shield of lighter materials that protects operators from γ rays and X-rays
In addition, reactors are often covered with a steel or concrete dome that is designed to contain any radioactive materials might be released by a
reactor accident.

6.8.4 https://chem.libretexts.org/@go/page/83458
I'm not a robot
reCAPTCHA
Privacy - Terms

About this page

Our systems have detected unusual traffic from your computer


network. This page checks to see if it's really you sending the
requests, and not a robot. Why did this happen?

IP address: 2604:a880:2:d0::2056:8001
Time: 2023-03-12T20:33:03Z
URL: https://www.youtube-nocookie.com/embed/jMFdo0n1Nto?
vq=hd1080

Video 6.8.1 : Click here to watch a 3-minute video from the Nuclear Energy Institute on how nuclear reactors work.
Nuclear power plants are designed in such a way that they cannot form a supercritical mass of fissionable material and therefore cannot create a
nuclear explosion. But as history has shown, failures of systems and safeguards can cause catastrophic accidents, including chemical explosions and
nuclear meltdowns (damage to the reactor core from overheating). The following Chemistry in Everyday Life feature explores three infamous
meltdown incidents.

Nuclear Accidents
The importance of cooling and containment are amply illustrated by three major accidents that occurred with the nuclear reactors at nuclear power
generating stations in the United States (Three Mile Island), the former Soviet Union (Chernobyl), and Japan (Fukushima).
In March 1979, the cooling system of the Unit 2 reactor at Three Mile Island Nuclear Generating Station in Pennsylvania failed, and the cooling
water spilled from the reactor onto the floor of the containment building. After the pumps stopped, the reactors overheated due to the high
radioactive decay heat produced in the first few days after the nuclear reactor shut down. The temperature of the core climbed to at least 2200 °C,
and the upper portion of the core began to melt. In addition, the zirconium alloy cladding of the fuel rods began to react with steam and produced
hydrogen:
Zr(s) + 2 H O(g) ⟶ ZrO (s) + 2 H (g) (6.8.2)
2 2 2

The hydrogen accumulated in the confinement building, and it was feared that there was danger of an explosion of the mixture of hydrogen and air
in the building. Consequently, hydrogen gas and radioactive gases (primarily krypton and xenon) were vented from the building. Within a week,
cooling water circulation was restored and the core began to cool. The plant was closed for nearly 10 years during the cleanup process.
Although zero discharge of radioactive material is desirable, the discharge of radioactive krypton and xenon, such as occurred at the Three Mile
Island plant, is among the most tolerable. These gases readily disperse in the atmosphere and thus do not produce highly radioactive areas.
Moreover, they are noble gases and are not incorporated into plant and animal matter in the food chain. Effectively none of the heavy elements of the
core of the reactor were released into the environment, and no cleanup of the area outside of the containment building was necessary (Figure 6.8.8).

Figure 6.8.8 : (a) In this 2010 photo of Three Mile Island, the remaining structures from the damaged Unit 2 reactor are seen on the left, whereas the
separate Unit 1 reactor, unaffected by the accident, continues generating power to this day (right). (b) President Jimmy Carter visited the Unit 2
control room a few days after the accident in 1979.

6.8.5 https://chem.libretexts.org/@go/page/83458
Another major nuclear accident involving a reactor occurred in April 1986, at the Chernobyl Nuclear Power Plant in Ukraine, which was still a part
of the former Soviet Union. While operating at low power during an unauthorized experiment with some of its safety devices shut off, one of the
reactors at the plant became unstable. Its chain reaction became uncontrollable and increased to a level far beyond what the reactor was designed for.
The steam pressure in the reactor rose to between 100 and 500 times the full power pressure and ruptured the reactor. Because the reactor was not
enclosed in a containment building, a large amount of radioactive material spewed out, and additional fission products were released, as the graphite
(carbon) moderator of the core ignited and burned. The fire was controlled, but over 200 plant workers and firefighters developed acute radiation
sickness and at least 32 soon died from the effects of the radiation. It is predicted that about 4000 more deaths will occur among emergency workers
and former Chernobyl residents from radiation-induced cancer and leukemia. The reactor has since been encapsulated in steel and concrete, a now-
decaying structure known as the sarcophagus. Almost 30 years later, significant radiation problems still persist in the area, and Chernobyl largely
remains a wasteland.
In 2011, the Fukushima Daiichi Nuclear Power Plant in Japan was badly damaged by a 9.0-magnitude earthquake and resulting tsunami. Three
reactors up and running at the time were shut down automatically, and emergency generators came online to power electronics and coolant systems.
However, the tsunami quickly flooded the emergency generators and cut power to the pumps that circulated coolant water through the reactors.
High-temperature steam in the reactors reacted with zirconium alloy to produce hydrogen gas. The gas escaped into the containment building, and
the mixture of hydrogen and air exploded. Radioactive material was released from the containment vessels as the result of deliberate venting to
reduce the hydrogen pressure, deliberate discharge of coolant water into the sea, and accidental or uncontrolled events.
An evacuation zone around the damaged plant extended over 12.4 miles away, and an estimated 200,000 people were evacuated from the area. All
48 of Japan’s nuclear power plants were subsequently shut down, remaining shuttered as of December 2014. Since the disaster, public opinion has
shifted from largely favoring to largely opposing increasing the use of nuclear power plants, and a restart of Japan’s atomic energy program is still
stalled (Figure 6.8.10).

Figure 6.8.9 : (a) After the accident, contaminated waste had to be removed, and (b) an evacuation zone was set up around the plant in areas that
received heavy doses of radioactive fallout. (credit a: modification of work by “Live Action Hero”/Flickr)
The energy produced by a reactor fueled with enriched uranium results from the fission of uranium as well as from the fission of plutonium
produced as the reactor operates. As discussed previously, the plutonium forms from the combination of neutrons and the uranium in the fuel. In any
nuclear reactor, only about 0.1% of the mass of the fuel is converted into energy. The other 99.9% remains in the fuel rods as fission products and
unused fuel. All of the fission products absorb neutrons, and after a period of several months to a few years, depending on the reactor, the fission
products must be removed by changing the fuel rods. Otherwise, the concentration of these fission products would increase and absorb more
neutrons until the reactor could no longer operate.
Spent fuel rods contain a variety of products, consisting of unstable nuclei ranging in atomic number from 25 to 60, some transuranium elements,
including plutonium and americium, and unreacted uranium isotopes. The unstable nuclei and the transuranium isotopes give the spent fuel a
dangerously high level of radioactivity. The long-lived isotopes require thousands of years to decay to a safe level. The ultimate fate of the nuclear
reactor as a significant source of energy in the United States probably rests on whether or not a politically and scientifically satisfactory technique
for processing and storing the components of spent fuel rods can be developed.

Contributors and Attributions


Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley (Stephen F. Austin
State University) with contributing authors. Textbook content produced by OpenStax College is licensed under a Creative Commons Attribution
License 4.0 license. Download for free at http://cnx.org/contents/[email protected]).

6.8: Nuclear Fission is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

6.8.6 https://chem.libretexts.org/@go/page/83458
6.9: Nuclear Fusion
Skills to Develop
Describe the nuclear reactions in a nuclear fusion reaction
Quantify the energy released or absorbed in a fusion reaction

The process of converting very light nuclei into heavier nuclei is also accompanied by the conversion of mass into large amounts of
energy, a process called fusion. The principal source of energy in the sun is a net fusion reaction in which four hydrogen nuclei fuse
and produce one helium nucleus and two positrons. This is a net reaction of a more complicated series of events:
1 4 0 +
4 H ⟶ 2
He + 2 β (6.9.1)
1 1
  
4×1.00794 u 4.0012 u 2×0.0005487 u

A helium nucleus has a mass that is 0.7% less than that of four hydrogen nuclei; this lost mass is converted into thermal energy
during the fusion via Einstein's E = mc equation. This reaction produces about 3.6 × 1011 kJ of energy per mole of He
2 4
2

produced. This is somewhat larger than the energy produced by the nuclear fission of one mole of U (1.8 × 1010 kJ), and over 3
235

million times larger than the energy produced by the (chemical) combustion of one mole of octane (ΔH = −5471 kJ ):
2 C8 H18 + 25 O2 → 16C O2 + 18 H2 O (6.9.2)

It has been determined that the nuclei of the heavy isotopes of hydrogen, a deuteron, H and a triton, 2
1
3
1
H , undergo fusion at
extremely high temperatures (thermonuclear fusion). They form a helium nucleus and a neutron:
2 3 4 1
1
H + 1H ⟶ 2
He + 2 n (6.9.3)
0

This change proceeds with a mass loss of 0.0188 amu, corresponding to the release of 1.69 × 109 kilojoules per mole of He 4
2

formed. The very high temperature is necessary to give the nuclei enough kinetic energy to overcome the very strong repulsive
forces resulting from the positive charges on their nuclei so they can collide.

Figure 6.9.1 : Fusion of deuterium with tritium creating helium-4, freeing a neutron, and releasing 17.59 MeV of energy, as an
appropriate amount of mass changing forms to appear as the kinetic energy of the products, in agreement with kinetic E = Δmc , 2

where Δm is the change in rest mass of particles.[Image use with permission via Wikipedia (Wykis)
The most important fusion process in nature is the one that powers stars. In the 20th century, it was realized that the energy released
from nuclear fusion reactions accounted for the longevity of the Sun and other stars as a source of heat and light. The fusion of
nuclei in a star, starting from its initial hydrogen and helium abundance, provides that energy and synthesizes new nuclei as a
byproduct of that fusion process. The prime energy producer in the Sun is the fusion of hydrogen to form helium, which occurs at a
solar-core temperature of 14 million kelvin. The net result is the fusion of four protons into one alpha particle, with the release of
two positrons, two neutrinos (which changes two of the protons into neutrons), and energy (Figure 6.9.2).

6.9.1 https://chem.libretexts.org/@go/page/83459
Figure 6.9.2 : (left) The Sun is a main-sequence star, and thus generates its energy by nuclear fusion of hydrogen nuclei into
helium. In its core, the Sun fuses 620 million metric tons of hydrogen each second. (right) The proton-proton chain dominates in
stars the size of the Sun or smaller.

Example 6.9.1

Calculate the energy released in each of the following hypothetical processes.


a. 3 4
2
He →
12
6
C

b. 6 1
1
H+6
1
0
n →
12
6
C

c. 6 2
1
D →
12
6
C

Discuss your results.


a. Qa = 3 * 4.0026 - 12.000) amu * (1.4924E-10 J/amu)
= 1.17E-12 J
b. Qb = (6*(1.007825 + 1.008665) - 12.00000) amu * (1.4924E-10 J/amu)
= 1.476E-11 J
c. Qc = 6*2.014102 - 12.00000 amu * (1.4924E-10 J/amu)
= 1.263E-11 J
Fusion of He to give C releases the least amount of energy, because the fusion to produce He has released a large amount. The
difference between the second and the third is the binding energy of deuterium. The conservation of mass-and-energy is well
illustrated in these calculations. On the other hand, the calculation is based on the conservation of mass-and-energy.

Nuclear Reactors
Useful fusion reactions require very high temperatures for their initiation—about 15,000,000 K or more. At these temperatures, all
molecules dissociate into atoms, and the atoms ionize, forming plasma. These conditions occur in an extremely large number of
locations throughout the universe—stars are powered by fusion. Humans have already figured out how to create temperatures high
enough to achieve fusion on a large scale in thermonuclear weapons. A thermonuclear weapon such as a hydrogen bomb contains a
nuclear fission bomb that, when exploded, gives off enough energy to produce the extremely high temperatures necessary for
fusion to occur.

6.9.2 https://chem.libretexts.org/@go/page/83459
Figure 6.9.3 : (a) This model is of the International Thermonuclear Experimental Reactor (ITER) reactor. Currently under
construction in the south of France with an expected completion date of 2027, the ITER will be the world’s largest experimental
Tokamak nuclear fusion reactor with a goal of achieving large-scale sustained energy production. (b) In 2012, the National Ignition
Facility at Lawrence Livermore National Laboratory briefly produced over 500,000,000,000 watts (500 terawatts, or 500 TW) of
peak power and delivered 1,850,000 joules (1.85 MJ) of energy, the largest laser energy ever produced and 1000 times the power
usage of the entire United States in any given moment. Although lasting only a few billionths of a second, the 192 lasers attained
the conditions needed for nuclear fusion ignition. This image shows the target prior to the laser shot. (credit a: modification of work
by Stephan Mosel)
Another much more beneficial way to create fusion reactions is in a fusion reactor, a nuclear reactor in which fusion reactions of
light nuclei are controlled. Because no solid materials are stable at such high temperatures, mechanical devices cannot contain the
plasma in which fusion reactions occur. Two techniques to contain plasma at the density and temperature necessary for a fusion
reaction are currently the focus of intensive research efforts: containment by a magnetic field and by the use of focused laser beams
(Figure 6.9.3). A number of large projects are working to attain one of the biggest goals in science: getting hydrogen fuel to ignite
and produce more energy than the amount supplied to achieve the extremely high temperatures and pressures that are required for
fusion. At the time of this writing, there are no self-sustaining fusion reactors operating in the world, although small-scale
controlled fusion reactions have been run for very brief periods.

Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).

6.9: Nuclear Fusion is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

6.9.3 https://chem.libretexts.org/@go/page/83459
6.10: Nuclear Chemistry (Exercises)

A general chemistry Libretexts Textmap organized around the textbook

 Chemistry: The Central Science


by Brown, LeMay, Bursten, Murphy, and Woodward

These are homework exercises to accompany the Textmap created for "Chemistry: The Central Science" by Brown et al.
Complementary General Chemistry question banks can be found for other Textmaps and can be accessed here. In addition to these
publicly available questions, access to private problems bank for use in exams and homework is available to faculty only on an
individual basis; please contact Delmar Larsen for an account with access permission.

21.1: Radioactivity
Q21.1.1
Why are many radioactive substances warm to the touch? Why do many radioactive substances glow?

Q21.1.2
Describe the differences between nonionizing and ionizing radiation in terms of the intensity of energy emitted and the effect each
has on an atom or molecule after collision. Which nuclear decay reactions are more likely to produce ionizing radiation?
nonionizing radiation?

Q21.1.3
Would you expect nonionizing or ionizing radiation to be more effective at treating cancer? Why?

S21.1.3
Ionizing radiation is higher in energy and causes greater tissue damage, so it is more likely to destroy cancerous cells.

Q21.1.4
Historically, concrete shelters have been used to protect people from nuclear blasts. Comment on the effectiveness of such shelters.

Q21.1.5
Gamma rays are a very high-energy radiation, yet α particles inflict more damage on biological tissue. Why?

Q21.1.6
List the three primary sources of naturally occurring radiation. Explain the factors that influence the dose that one receives
throughout the year. Which is the largest contributor to overall exposure? Which is the most hazardous?

S21.1.6
Three primary naturally occurring radiations are radium, uranium and thorium, each all having long half lives. Inhalation of air,
ingestion of food and water,terrestrial radation from the ground and cosmic radiation from space are all factors tat influence the
does that a person receives throughout the year. Inhalation of the air is the largest contributor to exposure. Radiation can damage
DNA or kill cells. When radiation is exposed to your body, it will collide with atoms and this will change and damage your DNA.

Q21.1.7
Because radon is a noble gas, it is inert and generally unreactive. Despite this, exposure to even low concentrations of radon in air
is quite dangerous. Describe the physical consequences of exposure to radon gas. Why are people who smoke more susceptible to
these effects?

6.10.1 https://chem.libretexts.org/@go/page/6498
Q21.1.8
Most medical imaging uses isotopes that have extremely short half-lives. These isotopes usually undergo only one kind of nuclear
decay reaction. Which kind of decay reaction is usually used? Why? Why would a short half-life be preferred in these cases?

Beta decay. Alfa decay can be easily stopped by paper, which means it can not be used to see inside people's body. Also, Gamma
rays are really dangerous for human, that even a short period of time exploding to it will have negative effect on human body. Thus,
Beta decay is the perfect choice. It can be used to see through human's body and stopped by aluminum or some other metals.

Since all these radioactive decays are harmful for human body, if the half time of these reactions are short, the time exploding to
these reactions will be short too.

Q21.1.9
Which would you prefer: one exposure of 100 rem, or 10 exposures of 10 rem each? Explain your rationale.

S21.1.9
Ten exposures of 10 rem are less likely to cause major damage.

Q21.1.10
A 2.14 kg sample of rock contains 0.0985 g of uranium. How much energy is emitted over 25 yr if 99.27% of the uranium is 238U,
which has a half-life of 4.46 × 109 yr, if each decay event is accompanied by the release of 4.039 MeV? If a 180 lb individual
absorbs all of the emitted radiation, how much radiation has been absorbed in rads?

Q21.1.11
There is a story about a “radioactive boy scout” who attempted to convert thorium-232, which he isolated from about 1000 gas
lantern mantles, to uranium-233 by bombarding the thorium with neutrons. The neutrons were generated via bombarding an
aluminum target with α particles from the decay of americium-241, which was isolated from 100 smoke detectors. Write balanced
nuclear reactions for these processes. The “radioactive boy scout” spent approximately 2 h/day with his experiment for 2 yr.
Assuming that the alpha emission of americium has an energy of 5.24 MeV/particle and that the americium-241 was undergoing
3.5 × 106 decays/s, what was the exposure of the 60.0 kg scout in rads? The intrepid scientist apparently showed no ill effects from
this exposure. Why?

241/95 Am---> 4/2 He + 237/93Np---> 4/2He + 233/91Pa----> 1/0n+ 232/91Th---> 1/1 H + 233/92 U

By adding alpha particles to the products side of the reaction, he was able to reduce the mass number by 4 and the atomic number
by 2 to get the products he wanted. Bombardment with neutrons and 1 H was required to lower to the mass number to get Th and
then raise both the mass number and the atomic number to yield Uranium.

2 hours*365*2= 1460 hours of exposure.*60min/1hr*60s/1min= 5.26*10^6s of exposure

1MeV= (1.6022*10^-13 Joules) * (5.24 MeV/particle)*2 particles= 1.679*10^-12 Joules.

(1.679*10^-12 Joules) * (1 amu/ 1.4924*10^-10 Joules)= 2.51*10^-13 amu

E=mc^2

E=(2.51*10^-13 amu)(1.66*10^-22kg/amu)(2.9998*10^8m/s)^2= (3.75*10^-18 kgm^2/s)*(3.5*10^6 decays/s)= 1.31*10^-11


joules of exposure per second.

The scientist showed no ill effects from this exposure because if we multiple the energy in joules of exposure per second,
1.31*10^-11, by the total amount of seconds of exposure, 5.26*10^6s, we find that he was only exposed to 6.9*10^-5 joules of
radiation throughout the span of two years. This is a very small amount of radiation for such a long span of time.

6.10.2 https://chem.libretexts.org/@go/page/6498
In order to plug in the values for this equation, we must convert the given MeV to Joules with the known conversion rate. Similarly,
we must convert Joules to amu with another known conversion rate. Then we can plug in the values and multiply by c^2 but we
must not forget to multiple the amu by the conversion rate to kg in order to yield Joules. After all of this is done, we multiple the
amount of Joules of exposure per second by the total amount of exposure in seconds in order to find out the total amount of
exposure over the two year span.

21.2: Patterns of Nuclear Stability


Q21.2.1
How do chemical reactions compare with nuclear reactions with respect to mass changes? Does either type of reaction violate the
law of conservation of mass? Explain your answers.

Q21.2.2
Why is the amount of energy released by a nuclear reaction so much greater than the amount of energy released by a chemical
reaction?

Q21.2.3
Explain why the mass of an atom is less than the sum of the masses of its component particles.

Q21.2.4
The stability of a nucleus can be described using two values. What are they, and how do they differ from each other?

Q21.2.5
In the days before true chemistry, ancient scholars (alchemists) attempted to find the philosopher’s stone, a material that would
enable them to turn lead into gold. Is the conversion of Pb → Au energetically favorable? Explain why or why not.

Q21.2.6
Describe the energy barrier to nuclear fusion reactions and explain how it can be overcome.

Q21.2.7
Imagine that the universe is dying, the stars have burned out, and all the elements have undergone fusion or radioactive decay.
What would be the most abundant element in this future universe? Why?

Q21.2.8
Numerous elements can undergo fission, but only a few can be used as fuels in a reactor. What aspect of nuclear fission allows a
nuclear chain reaction to occur?

Q21.2.9
How are transmutation reactions and fusion reactions related? Describe the main impediment to fusion reactions and suggest one or
two ways to surmount this difficulty.

Q21.2.10
Using the information provided in Chapter 33, complete each reaction and calculate the amount of energy released from each in
kilojoules.
a. 238Pa → ? + β−
b. 216Fr → ? + α
c. 199Bi → ? + β+

Q21.2.22
Using the information provided in Chapter 33, complete each reaction and calculate the amount of energy released from each in
kilojoules.
a. 194Tl → ? + β+

6.10.3 https://chem.libretexts.org/@go/page/6498
b. 171Pt → ? + α
c. 214Pb → ? + β−

Q21.2.23
Using the information provided in Chapter 33, complete each reaction and calculate the amount of energy released from each in
kilojoules per mole.
a. 234
91
Pa → ? +
0
−1
β

b. 226
88
Ra → ? +
4
2
α

Q21.2.24
Using the information provided in Chapter 33, complete each reaction and then calculate the amount of energy released from each
in kilojoules per mole.
a. Co → ? + β (The mass of cobalt-60 is 59.933817 amu.)
60
27
0
−1

b. technicium-94 (mass = 93.909657 amu) undergoing fission to produce chromium-52 and potassium-40

Q21.2.25
Using the information provided in Chapter 33, predict whether each reaction is favorable and the amount of energy released or
required in megaelectronvolts and kilojoules per mole.
a. the beta decay of bismuth-208 (mass = 207.979742 amu)
b. the formation of lead-206 by alpha decay

Q21.2.26
Using the information provided, predict whether each reaction is favorable and the amount of energy released or required in
megaelectronvolts and kilojoules per mole.
a. alpha decay of oxygen-16
b. alpha decay to produce chromium-52

Q21.2.27
Calculate the total nuclear binding energy (in megaelectronvolts) and the binding energy per nucleon for 87Sr if the measured mass
of 87Sr is 86.908877 amu.
a. the calculated mass
b. the mass defect
c. the nuclear binding energy
d. the nuclear binding energy per nucleon

Q21.2.30
The experimentally determined mass of 29S is 28.996610 amu. Find each of the following.
a. the calculated mass
b. the mass defect
c. the nuclear binding energy
d. the nuclear binding energy per nucleon

Q21.2.31
Calculate the amount of energy that is released by the neutron-induced fission of 235U to give 141Ba, 92Kr (mass = 91.926156 amu),
and three neutrons. Report your answer in electronvolts per atom and kilojoules per mole.

Q21.2.31
235 90 143
Calculate the amount of energy that is released by the neutron-induced fission of U to give Sr, Xe, and three neutrons.
Report your answer in electronvolts per atom and kilojoules per mole.

6.10.4 https://chem.libretexts.org/@go/page/6498
Q21.2.33
Calculate the amount of energy released or required by the fusion of helium-4 to produce the unstable beryllium-8 (mass =
8.00530510 amu). Report your answer in kilojoules per mole. Do you expect this to be a spontaneous reaction?

Q21.2.34
Calculate the amount of energy released by the fusion of 6Li and deuterium to give two helium-4 nuclei. Express your answer in
electronvolts per atom and kilojoules per mole.

Q21.2.35
How much energy is released by the fusion of two deuterium nuclei to give one tritium nucleus and one proton? How does this
amount compare with the energy released by the fusion of a deuterium nucleus and a tritium nucleus, which is accompanied by
ejection of a neutron? Express your answer in megaelectronvolts and kilojoules per mole. Pound for pound, which is a better choice
for a fusion reactor fuel mixture?

Numerical Answers
1
a. 238
91
Pa →
238
92
U+
0
−1
; −5.540 × 10−16 kJ
β
−15
b. 216
87
Fr →
212
85
At +
4
2
α ; −1.470 × 10 kJ
−16
c. 199
83
Bi →
199
82
Pb +
0
+1
β ; −5.458 × 10 kJ
3.
a. 234
91
Pa →
234
92
U+
0
−1
β ; 2.118 × 108 kJ/mol
8
b. 226
88
Ra →
222
86
Rn +
4
2
α ; 4.700 × 10 kJ/mol

5.
a. The beta decay of bismuth-208 to polonium is endothermic (ΔE = 1.400 MeV/atom, 1.352 × 108 kJ/mol).
b. The formation of lead-206 by alpha decay of polonium-210 is exothermic (ΔE = −5.405 MeV/atom, −5.218 × 108
kJ/mol).
7. 757 MeV/atom, 8.70 MeV/nucleon
9.
a. 53.438245 amu
b. 0.496955 amu
c. 463 MeV/atom
d. 8.74 MeV/nucleon
11. −173 MeV/atom; 1.67 × 1010 kJ/mol
13. ΔE = + 9.0 × 106 kJ/mol beryllium-8; no
15. D–D fusion: ΔE = −4.03 MeV/tritium nucleus formed = −3.89 × 108 kJ/mol tritium; D–T fusion: ΔE = −17.6 MeV/tritium
nucleus = −1.70 × 109 kJ/mol; D–T fusion

21.3: Nuclear Transmutations


Q21.3.1
Why do scientists believe that hydrogen is the building block of all other elements? Why do scientists believe that helium-4 is the
building block of the heavier elements?

Q21.3.2
How does a star produce such enormous amounts of heat and light? How are elements heavier than Ni formed?

Q21.3.3
Propose an explanation for the observation that elements with even atomic numbers are more abundant than elements with odd
atomic numbers.

6.10.5 https://chem.libretexts.org/@go/page/6498
S21.3.3
The raw material for all elements with Z > 2 is helium (Z = 2), and fusion of helium nuclei will always produce nuclei with an even
number of protons.

Q21.3.4
During the lifetime of a star, different reactions that form different elements are used to power the fusion furnace that keeps a star
“lit.” Explain the different reactions that dominate in the different stages of a star’s life cycle and their effect on the temperature of
the star.

Q21.3.5
A line in a popular song from the 1960s by Joni Mitchell stated, “We are stardust….” Does this statement have any merit or is it
just poetic? Justify your answer.

Q21.3.6
If the laws of physics were different and the primary element in the universe were boron-11 (Z = 5), what would be the next four
most abundant elements? Propose nuclear reactions for their formation.

Q21.3.7
Write a balanced nuclear reaction for the formation of each isotope.
a. 27Al from two 12C nuclei
b. 9Be from two 4He nuclei

Q21.3.8
At the end of a star’s life cycle, it can collapse, resulting in a supernova explosion that leads to the formation of heavy elements by
multiple neutron-capture events. Write a balanced nuclear reaction for the formation of each isotope during such an explosion.
a. 106Pd from nickel-58
b. selenium-79 from iron-56

Q21.3.9
When a star reaches middle age, helium-4 is converted to short-lived beryllium-8 (mass = 8.00530510 amu), which reacts with
another helium-4 to produce carbon-12. How much energy is released in each reaction (in megaelectronvolts)? How many atoms of
helium must be “burned” in this process to produce the same amount of energy obtained from the fusion of 1 mol of hydrogen
atoms to give deuterium?

21.4: Rates of Radioactive Decay


Q21.4.1
What do chemists mean by the half-life of a reaction?

Q21.4.2
If a sample of one isotope undergoes more disintegrations per second than the same number of atoms of another isotope, how do
their half-lives compare?

Q21.4.3
Half-lives for the reaction A + B → C were calculated at three values of [A]0, and [B] was the same in all cases. The data are listed
in the following table:

[A]0 (M) t½ (s)

0.50 420

0.75 280

1.0 210

6.10.6 https://chem.libretexts.org/@go/page/6498
Does this reaction follow first-order kinetics? On what do you base your answer?

S21.4.3
1. No; the reaction is second order in A because the half-life decreases with increasing reactant concentration according to t1/2 =
1/k[A0].

Q21.4.4
Ethyl-2-nitrobenzoate (NO2C6H4CO2C2H5) hydrolyzes under basic conditions. A plot of [NO2C6H4CO2C2H5] versus t was used to
calculate t½, with the following results:

[NO2C6H4CO2C2H5] (M/cm3) t½ (s)

0.050 240

0.040 300

0.030 400

Is this a first-order reaction? Explain your reasoning.

Q21.4.5
Azomethane (CH3N2CH3) decomposes at 600 K to C2H6 and N2. The decomposition is first order in azomethane. Calculate t½
from the data in the following table:

Time (s) P
CH3 N2 CH3
(atm)

0 8.2 × 10−2

2000 3.99 × 10−2

4000 1.94 × 10−2

How long will it take for the decomposition to be 99.9% complete?

S21.4.5
t1/2 = 1.92 × 103 s or 1920 s; 19100 s or 5.32 hrs.

Q21.4.6
The first-order decomposition of hydrogen peroxide has a half-life of 10.7 h at 20°C. What is the rate constant (expressed in s−1)
for this reaction? If you started with a solution that was 7.5 × 10−3 M H2O2, what would be the initial rate of decomposition (M/s)?
What would be the concentration of H2O2 after 3.3 h?

6.10: Nuclear Chemistry (Exercises) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.

6.10.7 https://chem.libretexts.org/@go/page/6498

You might also like