0% found this document useful (0 votes)
133 views

Full 4

Uploaded by

brie
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)
133 views

Full 4

Uploaded by

brie
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/ 348

BELLINGHAM

TECHNICAL COLLEGE
CHEM 121

Spencer Berger
Bellingham Technical College
Bellingham Technical College
Bellingham Technical College CHEM 121

Spencer Berger
This text is disseminated via the Open Education Resource (OER) LibreTexts Project (https://LibreTexts.org) and like the hundreds
of other texts available within this powerful platform, it is freely available for reading, printing and "consuming." Most, but not all,
pages in the library have licenses that may allow individuals to make changes, save, and print this book. Carefully
consult the applicable license(s) before pursuing such effects.
Instructors can adopt existing LibreTexts texts or Remix them to quickly build course-specific resources to meet the needs of their
students. Unlike traditional textbooks, LibreTexts’ web based origins allow powerful integration of advanced features and new
technologies to support learning.

The LibreTexts mission is to unite students, faculty and scholars in a cooperative effort to develop an easy-to-use online platform
for the construction, customization, and dissemination of OER content to reduce the burdens of unreasonable textbook costs to our
students and society. The LibreTexts project is a multi-institutional collaborative venture to develop the next generation of open-
access texts to improve postsecondary education at all levels of higher learning by developing an Open Access Resource
environment. The project currently consists of 14 independently operating and interconnected libraries that are constantly being
optimized by students, faculty, and outside experts to supplant conventional paper-based books. These free textbook alternatives are
organized within a central environment that is both vertically (from advance to basic level) and horizontally (across different fields)
integrated.
The LibreTexts libraries are Powered by NICE CXOne and are supported by the Department of Education Open Textbook Pilot
Project, the UC Davis Office of the Provost, the UC Davis Library, the California State University Affordable Learning Solutions
Program, and Merlot. This material is based upon work supported by the National Science Foundation under Grant No. 1246120,
1525057, and 1413739.
Any opinions, findings, and conclusions or recommendations expressed in this material are those of the author(s) and do not
necessarily reflect the views of the National Science Foundation nor the US Department of Education.
Have questions or comments? For information about adoptions or adaptions contact info@LibreTexts.org. 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 01/11/2024
TABLE OF CONTENTS
Licensing

1: Module 1 - Matter and Measurement


1.1: Chemicals Compose Ordinary Things
1.2: The Scope of Chemistry
1.3: What is Matter?
1.4: Classifying Matter According to Its State—Solid, Liquid, and Gas
1.5: Classifying Matter According to Its State—Solid, Liquid, and Gas
1.6: Classifying Matter According to Its Composition
1.7: Differences in Matter- Physical and Chemical Properties
1.8: Changes in Matter - Physical and Chemical Changes
1.9: Conservation of Mass - There is No New Matter
1.10: Compounds Display Constant Composition
1.11: Chemical Formulas - How to Represent Compounds
1.12: Energy
1.13: Energy and Chemical and Physical Change
1.14: Taking Measurements
1.15: Scientific Notation - Writing Large and Small Numbers
1.16: Significant Figures - Writing Numbers to Reflect Precision
1.17: Significant Figures in Calculations
1.18: The Basic Units of Measurement
1.19: Problem Solving and Unit Conversions
1.20: Solving Multi-step Conversion Problems
1.21: Units Raised to a Power
1.22: Density

2: Module 2 - Atoms and Bonding


2.1: Indivisible - The Atomic Theory
2.2: The Nuclear Atom
2.3: The Properties of Protons, Neutrons, and Electrons
2.4: Elements- Defined by Their Number of Protons
2.5: Elements- Defined by Their Number of Protons
2.6: Looking for Patterns - The Periodic Table
2.7: Ions - Losing and Gaining Electrons
2.8: Isotopes - When the Number of Neutrons Varies
2.9: Atomic Mass - The Average Mass of an Element’s Atoms
2.10: Light- Electromagnetic Radiation
2.11: The Electromagnetic Spectrum
2.12: The Bohr Model - Atoms with Orbits
2.13: The Quantum-Mechanical Model- Atoms with Orbitals
2.14: Quantum-Mechanical Orbitals and Electron Configurations
2.15: The Explanatory Power of the Quantum-Mechanical Model
2.16: Periodic Trends - Atomic Size, Ionization Energy, and Metallic Character
2.17: Electron Configurations and the Periodic Table
2.18: Representing Valence Electrons with Dots
2.19: Lewis Structures of Ionic Compounds- Electrons Transferred
2.20: Covalent Lewis Structures- Electrons Shared

1 https://chem.libretexts.org/@go/page/423706
2.21: Writing Lewis Structures for Covalent Compounds
2.22: Predicting the Shapes of Molecules
2.23: Electronegativity and Polarity - Why Oil and Water Do not Mix
2.24: A Molecular View of Elements and Compounds
2.25: Writing Formulas for Ionic Compounds
2.26: Nomenclature- Naming Compounds
2.27: Naming Ionic Compounds
2.28: Naming Molecular Compounds

3: Module 3 - Stoichiometry
3.1: Counting Nails by the Pound
3.2: Counting Atoms by the Gram
3.3: Counting Molecules by the Gram
3.4: Chemical Formulas as Conversion Factors
3.5: Mass Percent Composition of Compounds
3.6: Mass Percent Composition from a Chemical Formula
3.7: The Chemical Equation
3.8: How to Write Balanced Chemical Equations
3.9: Making Pancakes- Relationships Between Ingredients
3.10: Making Molecules- Mole-to-Mole Conversions
3.11: Making Molecules- Mass-to-Mass Conversions
3.12: Stoichiometry
3.13: Limiting Reactant and Theoretical Yield
3.14: Limiting Reactant, Theoretical Yield, and Percent Yield from Initial Masses of Reactants
3.15: Enthalpy Change is a Measure of the Heat Evolved or Absorbed

4: Module 4 - Gases and Liquids


4.1: Kinetic Molecular Theory- A Model for Gases
4.2: Pressure - The Result of Constant Molecular Collisions
4.3: Boyle’s Law - Pressure and Volume
4.4: Charles’s Law- Volume and Temperature
4.5: Gay-Lussac's Law- Temperature and Pressure
4.6: Avogadro’s Law- Volume and Moles
4.7: The Ideal Gas Law- Pressure, Volume, Temperature, and Moles
4.8: Mixtures of Gases - Why Deep-Sea Divers Breathe a Mixture of Helium and Oxygen
4.9: Interactions between Molecules
4.10: Properties of Liquids and Solids
4.11: Intermolecular Forces in Action- Surface Tension and Viscosity
4.12: Evaporation and Condensation
4.13: Melting, Freezing, and Sublimation
4.14: Intermolecular Forces- Dispersion, Dipole–Dipole, Hydrogen Bonding, and Ion-Dipole
4.15: Water - A Remarkable Molecule

5: Module 5 - Solution Chemistry


5.1: Solutions - Homogeneous Mixtures
5.2: Solutions of Solids Dissolved in Water- How to Make Rock Candy
5.3: Specifying Solution Concentration- Mass Percent
5.4: Specifying Solution Concentration- Molarity
5.5: Solution Dilution
5.6: Solution Stoichiometry
5.7: Electrolytes

2 https://chem.libretexts.org/@go/page/423706
5.8: Acids- Properties and Examples
5.9: Bases- Properties and Examples
5.10: Molecular Definitions of Acids and Bases
5.11: Reactions of Acids and Bases
5.12: Strong and Weak Acids and Bases
5.13: Acid–Base Titration
5.14: Water - Acid and Base in One
5.15: The pH and pOH Scales - Ways to Express Acidity and Basicity
5.16: Buffers- Solutions that Resist pH Change

Index
Glossary
Detailed Licensing
Detailed Licensing

3 https://chem.libretexts.org/@go/page/423706
Licensing
A detailed breakdown of this resource's licensing can be found in Back Matter/Detailed Licensing.

1 https://chem.libretexts.org/@go/page/423707
CHAPTER OVERVIEW

1: Module 1 - Matter and Measurement


1.1: Chemicals Compose Ordinary Things
1.2: The Scope of Chemistry
1.3: What is Matter?
1.4: Classifying Matter According to Its State—Solid, Liquid, and Gas
1.5: Classifying Matter According to Its State—Solid, Liquid, and Gas
1.6: Classifying Matter According to Its Composition
1.7: Differences in Matter- Physical and Chemical Properties
1.8: Changes in Matter - Physical and Chemical Changes
1.9: Conservation of Mass - There is No New Matter
1.10: Compounds Display Constant Composition
1.11: Chemical Formulas - How to Represent Compounds
1.12: Energy
1.13: Energy and Chemical and Physical Change
1.14: Taking Measurements
1.15: Scientific Notation - Writing Large and Small Numbers
1.16: Significant Figures - Writing Numbers to Reflect Precision
1.17: Significant Figures in Calculations
1.18: The Basic Units of Measurement
1.19: Problem Solving and Unit Conversions
1.20: Solving Multi-step Conversion Problems
1.21: Units Raised to a Power
1.22: Density

1: Module 1 - Matter and Measurement is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.

1
1.1: Chemicals Compose Ordinary Things
Chemistry is the branch of science dealing with the structure, composition, properties, and the reactive characteristics of matter.
Matter is anything that has mass and occupies space. Thus, chemistry is the study of literally everything around us—the liquids that
we drink, the gases we breathe, the composition of everything from the plastic case on your phone to the earth beneath your feet.
Moreover, chemistry is the study of the transformation of matter. Crude oil is transformed into more useful petroleum products,
such as gasoline and kerosene, by the process of refining. Some of these products are further transformed into plastics. Crude metal
ores are transformed into metals, that can then be fashioned into everything from foil to automobiles. Potential drugs are identified
from natural sources, isolated and then prepared in the laboratory. Their structures are systematically modified to produce the
pharmaceuticals that have led to vast advances in modern medicine. Chemistry is at the center of all of these processes; chemists
are the people that study the nature of matter and learn to design, predict, and control these chemical transformations. Within the
branches of chemistry you will find several apparent subdivisions. Inorganic chemistry, historically, focused on minerals and metals
found in the earth, while organic chemistry dealt with carbon-containing compounds that were first identified in living things.
Biochemistry is an outgrowth of the application of organic chemistry to biology and relates to the chemical basis for living things.
In the later chapters of this text we will explore organic and biochemistry in a bit more detail and you will notice examples of
organic compounds scattered throughout the text. Today, the lines between the various fields have blurred significantly and a
contemporary chemist is expected to have a broad background in all of these areas.
In this chapter, we will discuss some of the properties of matter and how chemists measure those properties. We will introduce
some of the vocabulary that is used throughout chemistry and the other physical sciences.
Let’s begin with matter. Matter is defined as any substance that has mass. It is important to distinguish here between weight and
mass. Weight is the result of the pull of gravity on an object. On the Moon, an object will weigh less than the same object on Earth
because the pull of gravity is less on the Moon. The mass of an object, however, is an inherent property of that object and does not
change, regardless of location, gravitational pull, or anything else. It is a property that is solely dependent on the quantity of matter
within the object.
Contemporary theories suggests that matter is composed of atoms. Atoms themselves are constructed from neutrons, protons and
electrons, along with an ever-increasing array of other subatomic particles. We will focus on the neutron, a particle having no
charge; the proton, which carries a positive charge; and the electron, which has a negative charge. Atoms are incredibly small. To
give you an idea of the size of an atom, a single copper penny contains approximately 28,000,000,000,000,000,000,000 atoms
(that’s 28 sextillion). Because atoms and subatomic particles are so small, their mass is not readily measured using pounds, ounces,
grams or any other scale that we would use on larger objects. Instead, the mass of atoms and subatomic particles is measured using
atomic mass units (abbreviated amu). The atomic mass unit is based on a scale that relates the mass of different types of atoms to
each other (using the most common form of the element carbon as a standard). The amu scale gives us a convenient means to
describe the masses of individual atoms and to do quantitative measurements concerning atoms and their reactions. Within an atom,
the neutron and proton both have a mass of one amu; the electron has a much smaller mass (about 0.0005 amu).

Figure 1.1.1 : Atoms are incredible small. To give you an idea of the size of an atom, a single copper penny contains approximately
28,000,000,000,000,000,000,000 atoms (that’s 28 sextillion).
Atomic theory places the neutron and the proton in the center of the atom in the nucleus. In an atom, the nucleus is very small, very
dense, carries a positive charge (from the protons) and contains virtually all of the mass of the atom. Electrons are placed in a
diffuse cloud surrounding the nucleus. The electron cloud carries a net negative charge (from the charge on the electrons) and in a
neutral atom there are always as many electrons in this cloud as there are protons in the nucleus (the positive charges in the nucleus
are balanced by the negative charges of the electrons, making the atom neutral).
An atom is characterized by the number of neutrons, protons and electrons that it possesses. Today, we recognize at least 116
different types of atoms, each type having a different number of protons in its nucleus. These different types of atoms are called

1.1.1 https://chem.libretexts.org/@go/page/423564
elements. The neutral element hydrogen (the lightest element) will always have one proton in its nucleus and one electron in the
cloud surrounding the nucleus. The element helium will always have two protons in its nucleus. It is the number of protons in the
nucleus of an atom that defines the identity of an element. Elements can, however, have differing numbers of neutrons in their
nucleus. For example, stable helium nuclei exist that contain one, or two neutrons (but they all have two protons). These different
types of helium atoms have different masses (3 or 4 amu) and they are called isotopes. For any given isotope, the sum of the
numbers of protons and neutrons in the nucleus is called the mass number. All elements exist as a collection of isotopes, and the
mass of an element that we use in chemistry, the atomic mass, is the average of the masses of these isotopes. For helium, there is
approximately one isotope of Helium-3 for every one million isotopes of Helium-4, hence the average atomic mass is very close to
4 (4.002602).
As different elements were discovered and named, abbreviations of their names were developed to allow for a convenient chemical
shorthand. The abbreviation for an element is called its chemical symbol. A chemical symbol consists of one or two letters, and the
relationship between the symbol and the name of the element is generally apparent. Thus helium has the chemical symbol He,
nitrogen is N, and lithium is Li. Sometimes the symbol is less apparent but is decipherable; magnesium is Mg, strontium is Sr, and
manganese is Mn. Symbols for elements that have been known since ancient times, however, are often based on Latin or Greek
names and appear somewhat obscure from their modern English names. For example, copper is Cu (from cuprum), silver is Ag
(from argentum), gold is Au (from aurum), and iron is Fe (from ferrum). Throughout your study of chemistry, you will routinely
use chemical symbols and it is important that you begin the process of learning the names and chemical symbols for the common
elements. By the time you complete General Chemistry, you will find that you are adept at naming and identifying virtually all of
the 116 known elements. Table 1.1.1 contains a starter list of common elements that you should begin learning now!
Table 1.1.1 : Names and Chemical Symbols for Common Elements
Element Chemical Symbol Element Chemical Symbol

Hydrogen H Phosphorus P

Helium He Sulfur S

Lithium Li Chlorine Cl

Beryllium Be Argon Ar

Boron B Potassium K

Carbon C Calcium Ca

Nitrogen N Iron Fe

Oxygen O Copper Cu

Fluorine F Zinc Zn

Neon Ne Bromine Br

Sodium Na Silver Ag

Magnesium Mg Iodine I

Aluminum Al Gold Au

Silicon Si Lead Pb

The chemical symbol for an element is often combined with information regarding the number of protons and neutrons in a
particular isotope of that atom to give the atomic symbol. To write an atomic symbol, begin with the chemical symbol, then write
the atomic number for the element (the number of protons in the nucleus) as a subscript, preceding the chemical symbol. Directly
above this, as a superscript, write the mass number for the isotope, that is, the total number of protons and neutrons in the nucleus.
Thus, for helium, the atomic number is 2 and there are two neutrons in the nucleus for the most common isotope, making the
atomic symbol He. In the definition of the atomic mass unit, the “most common isotope of carbon”, C, is defined as having a
4
2
12
6

mass of exactly 12 amu and the atomic masses of the remaining elements are based on their masses relative to this isotope.
Chlorine (chemical symbol Cl) consists of two major isotopes, one with 18 neutrons (the most common, comprising 75.77% of
natural chlorine atoms) and one with 20 neutrons (the remaining 24.23%). The atomic number of chlorine is 17 (it has 17 protons in
its nucleus), therefore the chemical symbols for the two isotopes are Cl and Cl .
35
17
37
17

1.1.2 https://chem.libretexts.org/@go/page/423564
When data is available regarding the natural abundance of various isotopes of an element, it is simple to calculate the average
atomic mass. In the example above, Cl was the most common isotope with an abundance of 75.77% and Cl had an abundance
35
17
37
17

of the remaining 24.23%. To calculate the average mass, first convert the percentages into fractions; that is, simply divide them by
100. Now, chlorine-35 represents a fraction of natural chlorine of 0.7577 and has a mass of 35 (the mass number). Multiplying
these, we get (0.7577 × 35) = 26.51. To this, we need to add the fraction representing chlorine-37, or (0.2423 × 37) = 8.965; adding,
(26.51 + 8.965) = 35.48, which is the weighted average atomic mass for chlorine. Whenever we do mass calculations involving
elements or compounds (combinations of elements), we always need to use average atomic masses.

Contributions & Attributions


Paul R. Young, Professor of Chemistry, University of Illinois at Chicago, Wiki: AskTheNerd; PRY askthenerd.com - pyoung
uic.edu; ChemistryOnline.com

1.1: Chemicals Compose Ordinary Things is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew &
Henry Agnew.
1.2: Chemicals Compose Ordinary Things by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

1.1.3 https://chem.libretexts.org/@go/page/423564
1.2: The Scope of Chemistry
 Learning Objectives
To recognize the breadth, depth, and scope of chemistry.
Define chemistry in relation to other sciences.
Identify the main disciplines of chemistry.

Chemistry is the study of matter—what it consists of, what its properties are, and how it changes. Matter is anything that has mass
and takes up space—that is, anything that is physically real. Some things are easily identified as matter—the screen on which you
are reading this book, for example. Others are not so obvious. Because we move so easily through air, we sometimes forget that it,
too, is matter. Because of this, chemistry is a science that has its fingers in just about everything. Being able to describe the
ingredients in a cake and how they change when the cake is baked, for example, is chemistry!
Chemistry is one branch of science. Science is the process by which we learn about the natural universe by observing, testing, and
then generating models that explain our observations. Because the physical universe is so vast, there are many different branches of
science (Figure 1.2.1). Thus, chemistry is the study of matter, biology is the study of living things, and geology is the study of
rocks and the earth. Mathematics is the language of science, and we will use it to communicate some of the ideas of chemistry.

Figure 1.2.1 : The Relationships between Some of the Major Branches of Science. Chemistry lies more or less in the middle, which
emphasizes its importance to many branches of science.
Although we divide science into different fields, there is much overlap among them. For example, some biologists and chemists
work in both fields so much that their work is called biochemistry. Similarly, geology and chemistry overlap in the field called
geochemistry. Figure 1.2.1 shows how many of the individual fields of science are related. At some level, all of these fields depend
on matter because they all involve "stuff"; because of this, chemistry has been called the "central science", linking them all
together.

There are many other fields of science, in addition to the ones (biology, medicine, etc.)
listed here.

 Example 1.2.1: Science Fields


Which fields of study are branches of science? Explain.
a. sculpture
b. astronomy

1.2.1 https://chem.libretexts.org/@go/page/423565
Solution
a. Sculpture is not considered a science because it is not a study of some aspect of the natural universe.
b. Astronomy is the study of stars and planets, which are part of the natural universe. Astronomy is therefore a field of
science.

 Exercise 1.2.1

Which fields of study are branches of science?


a. physiology (the study of the function of an animal’s or a plant’s body)
b. geophysics
c. agriculture
d. politics

Answer a:
yes
Answer b:
yes
Answer c:
yes
Answer d:
no

Areas of Chemistry
The study of modern chemistry has many branches, but can generally be broken down into five main disciplines, or areas of study:
Physical chemistry: Physical chemistry is the study of macroscopic properties, atomic properties, and phenomena in chemical
systems. A physical chemist may study such things as the rates of chemical reactions, the energy transfers that occur in
reactions, or the physical structure of materials at the molecular level.
Organic chemistry: Organic chemistry is the study of chemicals containing carbon. Carbon is one of the most abundant
elements on Earth and is capable of forming a tremendously vast number of chemicals (over twenty million so far). Most of the
chemicals found in all living organisms are based on carbon.
Inorganic chemistry: Inorganic chemistry is the study of chemicals that, in general, are not primarily based on carbon.
Inorganic chemicals are commonly found in rocks and minerals. One current important area of inorganic chemistry deals with
the design and properties of materials involved in energy and information technology.
Analytical chemistry: Analytical chemistry is the study of the composition of matter. It focuses on separating, identifying, and
quantifying chemicals in samples of matter. An analytical chemist may use complex instruments to analyze an unknown
material in order to determine its various components.
Biochemistry: Biochemistry is the study of chemical processes that occur in living things. Research may cover anything from
basic cellular processes up to understanding disease states so that better treatments can be developed.

Figure 1.2.2 : (left) Measurement of trace metals using atomic spectroscopy. (right) Measurement of hormone concentrations.
In practice, chemical research is often not limited to just one of the five major disciplines. A particular chemist may use
biochemistry to isolate a particular chemical found in the human body such as hemoglobin, the oxygen carrying component of red

1.2.2 https://chem.libretexts.org/@go/page/423565
blood cells. He or she may then proceed to analyze the hemoglobin using methods that would pertain to the areas of physical or
analytical chemistry. Many chemists specialize in areas that are combinations of the main disciplines, such as bioinorganic
chemistry or physical organic chemistry.

History of Chemistry
The history of chemistry is an interesting and challenging one. Very early chemists were often motivated mainly by the
achievement of a specific goal or product. Making perfume or soaps did not need a lot of theory, just a good recipe and careful
attention to detail. There was no standard way of naming materials (and no periodic table that we could all agree on). It is often
difficult to figure out exactly what a particular person was using. However, the science developed over the centuries by trial and
error.
Major progress was made toward putting chemistry on a solid foundation when Robert Boyle (1637-1691) began his research in
chemistry (Figure 1.2.3). He developed the basic ideas about the behavior of gases. He could then describe gases mathematically.
Boyle also helped form the idea that small particles could combine to form molecules. Many years later, John Dalton used these
ideas to develop the atomic theory.

Figure 1.2.3 : Robert Boyle.


The field of chemistry began to develop rapidly in the 1700's. Joseph Priestley (1733-1804) isolated and characterized several
gases: oxygen, carbon monoxide, and nitrous oxide. It was later discovered that nitrous oxide ("laughing gas") worked as an
anesthetic. This gas was used for that purpose for the first time in 1844 during a tooth extraction. Other gases discovered during
that time were chlorine, by C.W. Scheele (1742-1786) and nitrogen, by Antoine Lavoisier (1743-1794). Lavoisier has been
considered by many scholars to be the "father of chemistry". Among other accomplishments, he discovered the role of oxygen in
combustion and definitively formulated the law of conservation of matter.
Chemists continued to discover new compounds in the 1800's. The science also began to develop a more theoretical foundation.
John Dalton (1766-1844) put forth his atomic theory in 1807. This idea allowed scientists to think about chemistry in a much more
systematic way. Amadeo Avogadro (1776-1856) laid the groundwork for a more quantitative approach to chemistry by calculating
the number of particles in a given amount of a gas. A lot of effort was put forth in studying chemical reactions. These efforts led to
new materials being produced. Following the invention of the battery by Alessandro Volta (1745-1827), the field of
electrochemistry (both theoretical and applications) developed through major contributions by Humphry Davy (1778-1829) and
Michael Faraday (1791-1867). Other areas of the discipline also progressed rapidly.
It would take a large book to cover developments in chemistry during the twentieth century and up to today. One major area of
expansion was in the area of the chemistry of living processes. Research in photosynthesis in plants, the discovery and
characterization of enzymes as biochemical catalysts, elucidation of the structures of biomolecules such as insulin and DNA—these
efforts gave rise to an explosion of information in the field of biochemistry.
The practical aspects of chemistry were not ignored. The work of Volta, Davy, and Faraday eventually led to the development of
batteries that provided a source of electricity to power a number of devices (Figure 1.2.4).

1.2.3 https://chem.libretexts.org/@go/page/423565
Figure 1.2.4 : Battery developed by Volta. (CC BY-SA 3.0; (left) GuidoB and (right) Kkkdc).
Charles Goodyear (1800-1860) discovered the process of vulcanization, allowing a stable rubber product to be produced for the
tires of all the vehicles we have today. Louis Pasteur (1822-1895) pioneered the use of heat sterilization to eliminate unwanted
microorganisms in wine and milk. Alfred Nobel (1833-1896) invented dynamite (Figure 1.2.5). After his death, the fortune he
made from this product was used to fund the Nobel Prizes in science and the humanities. J.W. Hyatt (1837-1920) developed the
first plastic. Leo Baekeland (1863-1944) developed the first synthetic resin, widely used for inexpensive and sturdy dinnerware.

Figure 1.2.5 : Dynamite explosion in Panama, Central America (1908).


Today, chemistry continues to be essential to the development of new materials and technologies, from semiconductors for
electronics to powerful new medicines, and beyond.

Summary
Chemistry is the study of matter and the changes it undergoes and considers both macroscopic and microscopic information.
Matter is anything that has mass and occupies space.
The five main disciplines of chemistry are physical chemistry, organic chemistry, inorganic chemistry, analytical chemistry and
biochemistry.
Many civilizations contributed to the growth of chemistry. A lot of early chemical research focused on practical uses. Basic
chemistry theories were developed during the nineteenth century. New materials and batteries are a few of the products of
modern chemistry.

1.2: The Scope of Chemistry is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew & Henry
Agnew.
1.7: The Scope of Chemistry by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

1.2.4 https://chem.libretexts.org/@go/page/423565
1.3: What is Matter?
 Learning Objectives
Define matter and explain how it is composed of building blocks known as "atoms".

We are all familiar with matter. The definition of Matter is anything that has mass and volume (takes up space). For most common
objects that we deal with every day, it is fairly simple to demonstrate that they have mass and take up space. You might be able to
imagine, however, the difficulty for people several hundred years ago to demonstrate that air had mass and volume. Air (and all
other gases) are invisible to the eye, have very small masses compared to equal amounts of solids and liquids, and are quite easy to
compress (change volume). Without sensitive equipment, it would have been difficult to convince people that gases are matter.
Today, we can measure the mass of a small balloon when it is deflated and then blow it up, tie it off, and measure its mass again to
detect the additional mass due to the air inside. The mass of air, under room conditions, that occupies a one quart jar is
approximately 0.0002 pounds. This small amount of mass would have been difficult to measure in times before balances were
designed to accurately measure very small masses. Later, scientists were able to compress gases into such a small volume that the
gases turned into liquids, which made it clear that gases are matter.

Figure 1.3.1 : Everything from an ant, to a truck, to the earth, and even the entire galaxy is composed of matter. Images used with
permission from Wikipedia (CC_SA-BY-3.0; credit High Contrast).
Even though the universe consists of "things" as wildly different as ants and galaxies, the matter that makes up all of these "things"
is composed of a very limited number of building blocks. These building blocks are known as atoms, and so far, scientists have
discovered or created a grand total of 118 different types of atoms. Scientists have given a name to each different type of atom. A
substance that is composed of only one type of atom is called an element. At this point, what should amaze you is that all forms of
matter in our universe are made with only 118 different building blocks. In some ways, it's sort of like cooking a gourmet, five-
course meal using only three ingredients! How is it possible? To answer that question, you have to understand the ways in which
different elements are put together to form matter.
The most important method that nature uses to organize atoms into matter is the formation of molecules. Molecules are groups of
two or more atoms that have been bonded together. There are millions of different ways to bond atoms together, which means that
there are millions of different possible molecules. Each of these molecules has its own set of chemical properties, and it's these

1.3.1 https://chem.libretexts.org/@go/page/423566
properties with which chemists are most concerned. You will learn a lot more about atoms and molecules, including how they were
discovered, in a later part of the textbook.

Summary
All matter has mass and occupies space. All physical objects are made of matter. Matter itself is composed of tiny building blocks
known as "atoms". There are only 118 different types of atoms known to man. Frequently, atoms are bonded together to form
"molecules".

Contributions & Attributions


Wikipedia

1.3: What is Matter? is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew & Henry Agnew.
3.2: What is Matter? by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source: https://www.ck12.org/c/chemistry/.

1.3.2 https://chem.libretexts.org/@go/page/423566
1.4: Classifying Matter According to Its State—Solid, Liquid, and Gas
 Learning Objectives
To describe the solid, liquid and gas phases.

Water can take many forms. At low temperatures (below 0 C), it is a solid. When at "normal" temperatures (between 0 C and
o o

100 C), it is a liquid. While at temperatures above 100 C, water is a gas (steam). The state that water is in depends upon the
o o

temperature. Each state has its own unique set of physical properties. Matter typically exists in one of three states: solid, liquid, or
gas.

Figure 1.4.1 : Matter is usually classified into three classical states, with plasma sometimes added as a fourth state. From left to
right: quartz (solid), water (liquid), nitrogen dioxide (gas).
The state that a given substance exhibits is also a physical property. Some substances exist as gases at room temperature (oxygen
and carbon dioxide), while others, like water and mercury metal, exist as liquids. Most metals exist as solids at room temperature.
All substances can exist in any of these three states. Figure 1.4.2 shows the differences among solids, liquids, and gases at the
molecular level. A solid has definite volume and shape, a liquid has a definite volume but no definite shape, and a gas has neither a
definite volume nor shape.

Figure 1.4.2 : A Representation of the Solid, Liquid, and Gas States. (a) Solid O2 has a fixed volume and shape, and the molecules
are packed tightly together. (b) Liquid O2 conforms to the shape of its container but has a fixed volume; it contains relatively
densely packed molecules. (c) Gaseous O2 fills its container completely—regardless of the container’s size or shape—and consists
of widely separated molecules.

 Plasma: A Fourth State of Matter

Technically speaking, a fourth state of matter called plasma exists, but it does not naturally occur on earth, so we will omit it
from our study here.

A plasma globe operating in a darkened room. (CC BY-SA 3.0; Chocolateoak).

1.4.1 https://chem.libretexts.org/@go/page/423567
Solids
In the solid state, the individual particles of a substance are in fixed positions with respect to each other because there is not enough
thermal energy to overcome the intermolecular interactions between the particles. As a result, solids have a definite shape and
volume. Most solids are hard, but some (like waxes) are relatively soft. Many solids composed of ions can also be quite brittle.
Solids are defined by the following characteristics:
Definite shape (rigid)
Definite volume
Particles vibrate around fixed axes
If we were to cool liquid mercury to its freezing point of −39 C, and under the right pressure conditions, we would notice all of
o

the liquid particles would go into the solid state. Mercury can be solidified when its temperature is brought to its freezing point.
However, when returned to room temperature conditions, mercury does not exist in solid state for long, and returns back to its more
common liquid form.
Solids usually have their constituent particles arranged in a regular, three-dimensional array of alternating positive and negative
ions called a crystal. The effect of this regular arrangement of particles is sometimes visible macroscopically, as shown in Figure
1.4.3. Some solids, especially those composed of large molecules, cannot easily organize their particles in such regular crystals and

exist as amorphous (literally, “without form”) solids. Glass is one example of an amorphous solid.

Figure 1.4.3 : (left) The periodic crystalline lattice structure of quartz SiO in two-dimensions. (right) The random network
2

structure of glassy SiO in two-dimensions. Note that, as in the crystal, each Silicon atom is bonded to 4 oxygen atoms, where the
2

fourth oxygen atom is obscured from view in this plane. Images used with permission (public domain).

Liquids
If the particles of a substance have enough energy to partially overcome intermolecular interactions, then the particles can move
about each other while remaining in contact. This describes the liquid state. In a liquid, the particles are still in close contact, so
liquids have a definite volume. However, because the particles can move about each other rather freely, a liquid has no definite
shape and takes a shape dictated by its container.
Liquids have the following characteristics:
No definite shape (takes the shape of its container).
Has definite volume.
Particles are free to move over each other, but are still attracted to each other.
A familiar liquid is mercury metal. Mercury is an anomaly. It is the only metal we know of that is liquid at room temperature.
Mercury also has an ability to stick to itself (surface tension)—a property that all liquids exhibit. Mercury has a relatively high
surface tension, which makes it very unique. Here you see mercury in its common liquid form.

1.4.2 https://chem.libretexts.org/@go/page/423567
What does the LIQUID METAL Mercury l…
l…

Video 1.4.1 : Mercury boiling to become a gas.


If we heat liquid mercury to its boiling point of 357 C
o
under the right pressure conditions, we would notice all particles in the
liquid state go into the gas state.

Gases
If the particles of a substance have enough energy to completely overcome intermolecular interactions, then the particles can
separate from each other and move about randomly in space. This describes the gas state, which we will consider in more detail
elsewhere. Like liquids, gases have no definite shape, but unlike solids and liquids, gases have no definite volume either. The
change from solid to liquid usually does not significantly change the volume of a substance. However, the change from a liquid to a
gas significantly increases the volume of a substance, by a factor of 1,000 or more. Gases have the following characteristics:
No definite shape (takes the shape of its container)
No definite volume
Particles move in random motion with little or no attraction to each other
Highly compressible
Table 1.4.1 : Characteristics of the Three States of Matter
Characteristics Solids Liquids Gases

shape definite indefinite indefinite

volume definite definite indefinite

relative intermolecular interaction


strong moderate weak
strength

relative particle positions in contact and fixed in place in contact but not fixed not in contact, random positions

 Example 1.4.1

What state or states of matter does each statement, describe?


a. This state has a definite volume, but no definite shape.
b. This state has no definite volume.
c. This state allows the individual particles to move about while remaining in contact.

Solution
a. This statement describes the liquid state.
b. This statement describes the gas state.
c. This statement describes the liquid state.

1.4.3 https://chem.libretexts.org/@go/page/423567
 Exercise 1.4.1
What state or states of matter does each statement describe?
a. This state has individual particles in a fixed position with regard to each other.
b. This state has individual particles far apart from each other in space.
c. This state has a definite shape.

Answer a:
solid

Answer b:
gas

Answer c:
solid

Summary
Three states of matter exist—solid, liquid, and gas.
Solids have a definite shape and volume.
Liquids have a definite volume, but take the shape of the container.
Gases have no definite shape or volume.

1.4: Classifying Matter According to Its State—Solid, Liquid, and Gas is shared under a CK-12 license and was authored, remixed, and/or curated
by Marisa Alviar-Agnew & Henry Agnew.
3.3: Classifying Matter According to Its State—Solid, Liquid, and Gas by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12.
Original source: https://www.ck12.org/c/chemistry/.

1.4.4 https://chem.libretexts.org/@go/page/423567
1.5: Classifying Matter According to Its State—Solid, Liquid, and Gas
 Learning Objectives
To describe the solid, liquid and gas phases.

Water can take many forms. At low temperatures (below 0 C), it is a solid. When at "normal" temperatures (between 0 C and
o o

100 C), it is a liquid. While at temperatures above 100 C, water is a gas (steam). The state that water is in depends upon the
o o

temperature. Each state has its own unique set of physical properties. Matter typically exists in one of three states: solid, liquid, or
gas.

Figure 1.5.1 : Matter is usually classified into three classical states, with plasma sometimes added as a fourth state. From left to
right: quartz (solid), water (liquid), nitrogen dioxide (gas).
The state that a given substance exhibits is also a physical property. Some substances exist as gases at room temperature (oxygen
and carbon dioxide), while others, like water and mercury metal, exist as liquids. Most metals exist as solids at room temperature.
All substances can exist in any of these three states. Figure 1.5.2 shows the differences among solids, liquids, and gases at the
molecular level. A solid has definite volume and shape, a liquid has a definite volume but no definite shape, and a gas has neither a
definite volume nor shape.

Figure 1.5.2 : A Representation of the Solid, Liquid, and Gas States. (a) Solid O2 has a fixed volume and shape, and the molecules
are packed tightly together. (b) Liquid O2 conforms to the shape of its container but has a fixed volume; it contains relatively
densely packed molecules. (c) Gaseous O2 fills its container completely—regardless of the container’s size or shape—and consists
of widely separated molecules.

 Plasma: A Fourth State of Matter

Technically speaking, a fourth state of matter called plasma exists, but it does not naturally occur on earth, so we will omit it
from our study here.

A plasma globe operating in a darkened room. (CC BY-SA 3.0; Chocolateoak).

1.5.1 https://chem.libretexts.org/@go/page/423568
Solids
In the solid state, the individual particles of a substance are in fixed positions with respect to each other because there is not enough
thermal energy to overcome the intermolecular interactions between the particles. As a result, solids have a definite shape and
volume. Most solids are hard, but some (like waxes) are relatively soft. Many solids composed of ions can also be quite brittle.
Solids are defined by the following characteristics:
Definite shape (rigid)
Definite volume
Particles vibrate around fixed axes
If we were to cool liquid mercury to its freezing point of −39 C, and under the right pressure conditions, we would notice all of
o

the liquid particles would go into the solid state. Mercury can be solidified when its temperature is brought to its freezing point.
However, when returned to room temperature conditions, mercury does not exist in solid state for long, and returns back to its more
common liquid form.
Solids usually have their constituent particles arranged in a regular, three-dimensional array of alternating positive and negative
ions called a crystal. The effect of this regular arrangement of particles is sometimes visible macroscopically, as shown in Figure
1.5.3. Some solids, especially those composed of large molecules, cannot easily organize their particles in such regular crystals and

exist as amorphous (literally, “without form”) solids. Glass is one example of an amorphous solid.

Figure 1.5.3 : (left) The periodic crystalline lattice structure of quartz SiO in two-dimensions. (right) The random network
2

structure of glassy SiO in two-dimensions. Note that, as in the crystal, each Silicon atom is bonded to 4 oxygen atoms, where the
2

fourth oxygen atom is obscured from view in this plane. Images used with permission (public domain).

Liquids
If the particles of a substance have enough energy to partially overcome intermolecular interactions, then the particles can move
about each other while remaining in contact. This describes the liquid state. In a liquid, the particles are still in close contact, so
liquids have a definite volume. However, because the particles can move about each other rather freely, a liquid has no definite
shape and takes a shape dictated by its container.
Liquids have the following characteristics:
No definite shape (takes the shape of its container).
Has definite volume.
Particles are free to move over each other, but are still attracted to each other.
A familiar liquid is mercury metal. Mercury is an anomaly. It is the only metal we know of that is liquid at room temperature.
Mercury also has an ability to stick to itself (surface tension)—a property that all liquids exhibit. Mercury has a relatively high
surface tension, which makes it very unique. Here you see mercury in its common liquid form.

1.5.2 https://chem.libretexts.org/@go/page/423568
What does the LIQUID METAL Mercury l…
l…

Video 1.5.1 : Mercury boiling to become a gas.


If we heat liquid mercury to its boiling point of 357 C
o
under the right pressure conditions, we would notice all particles in the
liquid state go into the gas state.

Gases
If the particles of a substance have enough energy to completely overcome intermolecular interactions, then the particles can
separate from each other and move about randomly in space. This describes the gas state, which we will consider in more detail
elsewhere. Like liquids, gases have no definite shape, but unlike solids and liquids, gases have no definite volume either. The
change from solid to liquid usually does not significantly change the volume of a substance. However, the change from a liquid to a
gas significantly increases the volume of a substance, by a factor of 1,000 or more. Gases have the following characteristics:
No definite shape (takes the shape of its container)
No definite volume
Particles move in random motion with little or no attraction to each other
Highly compressible
Table 1.5.1 : Characteristics of the Three States of Matter
Characteristics Solids Liquids Gases

shape definite indefinite indefinite

volume definite definite indefinite

relative intermolecular interaction


strong moderate weak
strength

relative particle positions in contact and fixed in place in contact but not fixed not in contact, random positions

 Example 1.5.1

What state or states of matter does each statement, describe?


a. This state has a definite volume, but no definite shape.
b. This state has no definite volume.
c. This state allows the individual particles to move about while remaining in contact.

Solution
a. This statement describes the liquid state.
b. This statement describes the gas state.
c. This statement describes the liquid state.

1.5.3 https://chem.libretexts.org/@go/page/423568
 Exercise 1.5.1
What state or states of matter does each statement describe?
a. This state has individual particles in a fixed position with regard to each other.
b. This state has individual particles far apart from each other in space.
c. This state has a definite shape.

Answer a:
solid

Answer b:
gas

Answer c:
solid

Summary
Three states of matter exist—solid, liquid, and gas.
Solids have a definite shape and volume.
Liquids have a definite volume, but take the shape of the container.
Gases have no definite shape or volume.

1.5: Classifying Matter According to Its State—Solid, Liquid, and Gas is shared under a CK-12 license and was authored, remixed, and/or curated
by Marisa Alviar-Agnew & Henry Agnew.
3.3: Classifying Matter According to Its State—Solid, Liquid, and Gas by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12.
Original source: https://www.ck12.org/c/chemistry/.

1.5.4 https://chem.libretexts.org/@go/page/423568
1.6: Classifying Matter According to Its Composition
 Learning Objectives
Explain the difference between a pure substance and a mixture.
Explain the difference between an element and a compound.
Explain the difference between a homogeneous mixture and a heterogeneous mixture.

One useful way of organizing our understanding of matter is to think of a hierarchy that extends down from the most general and
complex to the simplest and most fundamental (Figure 1.6.1). Matter can be classified into two broad categories: pure substances
and mixtures. A pure substance is a form of matter that has a constant composition (meaning that it is the same everywhere) and
properties that are constant throughout the sample (meaning that there is only one set of properties such as melting point, color,
boiling point, etc. throughout the matter). A material composed of two or more substances is a mixture. Elements and compounds
are both examples of pure substances. A substance that cannot be broken down into chemically simpler components is an element.
Aluminum, which is used in soda cans, is an element. A substance that can be broken down into chemically simpler components
(because it has more than one element) is a compound. For example, water is a compound composed of the elements hydrogen and
oxygen. Today, there are about 118 elements in the known universe. In contrast, scientists have identified tens of millions of
different compounds to date.

Figure 1.6.1 : Relationships between the Types of Matter and the Methods Used to Separate Mixtures
Ordinary table salt is called sodium chloride. It is considered a substance because it has a uniform and definite composition. All
samples of sodium chloride are chemically identical. Water is also a pure substance. Salt easily dissolves in water, but salt water
cannot be classified as a substance because its composition can vary. You may dissolve a small amount of salt or a large amount
into a given amount of water. A mixture is a physical blend of two or more components, each of which retains its own identity and
properties in the mixture. Only the form of the salt is changed when it is dissolved into water. It retains its composition and
properties.

1.6.1 https://chem.libretexts.org/@go/page/423569
A homogeneous mixture is a mixture in which the composition is uniform throughout the mixture. The salt water described above
is homogeneous because the dissolved salt is evenly distributed throughout the entire salt water sample. Often it is easy to confuse
a homogeneous mixture with a pure substance because they are both uniform. The difference is that the composition of the
substance is always the same. The amount of salt in the salt water can vary from one sample to another. All solutions are
considered homogeneous because the dissolved material is present in the same amount throughout the solution.
A heterogeneous mixture is a mixture in which the composition is not uniform throughout the mixture. Vegetable soup is a
heterogeneous mixture. Any given spoonful of soup will contain varying amounts of the different vegetables and other components
of the soup.

 Phase

A phase is any part of a sample that has a uniform composition and properties. By definition, a pure substance or a
homogeneous mixture consists of a single phase. A heterogeneous mixture consists of two or more phases. When oil and water
are combined, they do not mix evenly, but instead form two separate layers. Each of the layers is called a phase.

 Example 1.6.1

Identify each substance as a compound, an element, a heterogeneous mixture, or a homogeneous mixture (solution).
a. filtered tea
b. freshly squeezed orange juice
c. a compact disc
d. aluminum oxide, a white powder that contains a 2:3 ratio of aluminum and oxygen atoms
e. selenium
Given: a chemical substance
Asked for: its classification

Strategy:
A. Decide whether a substance is chemically pure. If it is pure, the substance is either an element or a compound. If a
substance can be separated into its elements, it is a compound.
B. If a substance is not chemically pure, it is either a heterogeneous mixture or a homogeneous mixture. If its composition is
uniform throughout, it is a homogeneous mixture.

Solution
a. A) Tea is a solution of compounds in water, so it is not chemically pure. It is usually separated from tea leaves by filtration.
B) Because the composition of the solution is uniform throughout, it is a homogeneous mixture.
b. A) Orange juice contains particles of solid (pulp) as well as liquid; it is not chemically pure.
B) Because its composition is not uniform throughout, orange juice is a heterogeneous mixture.
c. A) A compact disc is a solid material that contains more than one element, with regions of different compositions visible
along its edge. Hence, a compact disc is not chemically pure.
B) The regions of different composition indicate that a compact disc is a heterogeneous mixture.
d. A) Aluminum oxide is a single, chemically pure compound.
e. A) Selenium is one of the known elements.

 Exercise 1.6.1

Identify each substance as a compound, an element, a heterogeneous mixture, or a homogeneous mixture (solution).
a. white wine
b. mercury
c. ranch-style salad dressing
d. table sugar (sucrose)

1.6.2 https://chem.libretexts.org/@go/page/423569
Answer a:
homogeneous mixture (solution)
Answer b:
element
Answer c:
heterogeneous mixture
Answer d:
compound

 Example 1.6.2

How would a chemist categorize each example of matter?


a. saltwater
b. soil
c. water
d. oxygen

Solution
a. Saltwater acts as if it were a single substance even though it contains two substances—salt and water. Saltwater is a
homogeneous mixture, or a solution.
b. Soil is composed of small pieces of a variety of materials, so it is a heterogeneous mixture.
c. Water is a substance. More specifically, because water is composed of hydrogen and oxygen, it is a compound.
d. Oxygen, a substance, is an element.

 Exercise 1.6.2

How would a chemist categorize each example of matter?


a. coffee
b. hydrogen
c. an egg

Answer a:
a homogeneous mixture (solution), assuming it is filtered coffee
Answer b:
element
Answer c:
heterogeneous mixture

Summary
Matter can be classified into two broad categories: pure substances and mixtures. A pure substance is a form of matter that has a
constant composition and properties that are constant throughout the sample. Mixtures are physical combinations of two or more
elements and/or compounds. Mixtures can be classified as homogeneous or heterogeneous. Elements and compounds are both
examples of pure substances. Compounds are substances that are made up of more than one type of atom. Elements are the simplest
substances made up of only one type of atom.

Vocabulary
Element: a substance that is made up of only one type of atom.
Compound:a substance that is made up of more than one type of atom bonded together.
Mixture: a combination of two or more elements or compounds which have not reacted to bond together; each part in the
mixture retains its own properties.

1.6.3 https://chem.libretexts.org/@go/page/423569
Contributions & Attributions

Stephen Lower, Professor Emeritus (Simon Fraser U.) Chem1 Virtual Textbook

1.6: Classifying Matter According to Its Composition is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa
Alviar-Agnew & Henry Agnew.
3.4: Classifying Matter According to Its Composition by Henry Agnew, Marisa Alviar-Agnew, Stephen Lower is licensed CK-12. Original
source: https://www.ck12.org/c/chemistry/.

1.6.4 https://chem.libretexts.org/@go/page/423569
1.7: Differences in Matter- Physical and Chemical Properties
 Learning Objectives

To separate physical from chemical properties.

All matter has physical and chemical properties. Physical properties are characteristics that scientists can measure without changing
the composition of the sample under study, such as mass, color, and volume (the amount of space occupied by a sample). Chemical
properties describe the characteristic ability of a substance to react to form new substances; they include its flammability and
susceptibility to corrosion. All samples of a pure substance have the same chemical and physical properties. For example, pure
copper is always a reddish-brown solid (a physical property) and always dissolves in dilute nitric acid to produce a blue solution
and a brown gas (a chemical property).

Physical Property
A physical property is a characteristic of a substance that can be observed or measured without changing the identity of the
substance. Silver is a shiny metal that conducts electricity very well. It can be molded into thin sheets, a property called
malleability. Salt is dull and brittle and conducts electricity when it has been dissolved into water, which it does quite easily.
Physical properties of matter include color, hardness, malleability, solubility, electrical conductivity, density, melting point, and
boiling point.
For the elements, color does not vary much from one element to the next. The vast majority of elements are colorless, silver, or
gray. Some elements do have distinctive colors: sulfur and chlorine are yellow, copper is (of course) copper-colored, and elemental
bromine is red. However, density can be a very useful parameter for identifying an element. Of the materials that exist as solids at
room temperature, iodine has a very low density compared to zinc, chromium, and tin. Gold has a very high density, as does
platinum. Pure water, for example, has a density of 0.998 g/cm3 at 25°C. The average densities of some common substances are in
Table 1.7.1. Notice that corn oil has a lower mass to volume ratio than water. This means that when added to water, corn oil will
“float.”
Table 1.7.1 : Densities of Common Substances
Substance Density at 25°C (g/cm3)

blood 1.035

body fat 0.918

whole milk 1.030

corn oil 0.922

mayonnaise 0.910

honey 1.420

Hardness helps determine how an element (especially a metal) might be used. Many elements are fairly soft (silver and gold, for
example) while others (such as titanium, tungsten, and chromium) are much harder. Carbon is an interesting example of hardness.
In graphite, (the "lead" found in pencils) the carbon is very soft, while the carbon in a diamond is roughly seven times as hard.

1.7.1 https://chem.libretexts.org/@go/page/423570
Figure 1.7.1 : Pencil (left) and Diamond ring (right). Both are a form of carbon, but exhibit very different physical properties.
Melting and boiling points are somewhat unique identifiers, especially of compounds. In addition to giving some idea as to the
identity of the compound, important information can be obtained about the purity of the material.

Chemical Properties
Chemical properties of matter describe its potential to undergo some chemical change or reaction by virtue of its composition.
The elements, electrons, and bonds that are present give the matter potential for chemical change. It is quite difficult to define a
chemical property without using the word "change". Eventually, after studying chemistry for some time, you should be able to look
at the formula of a compound and state some chemical property. For example, hydrogen has the potential to ignite and explode
given the right conditions—this is a chemical property. Metals in general have the chemical property of reacting with an acid. Zinc
reacts with hydrochloric acid to produce hydrogen gas—this is a chemical property.

Figure 1.7.2 : Heavy rust on the links of a chain near the Golden Gate Bridge in San Francisco; it was continuously exposed to
moisture and salt spray, causing surface breakdown, cracking, and flaking of the metal. (CC BY-SA 3.0; Marlith).
A chemical property of iron is its capability of combining with oxygen to form iron oxide, the chemical name of rust (Figure
1.7.2). The more general term for rusting and other similar processes is corrosion. Other terms that are commonly used in

descriptions of chemical changes are burn, rot, explode, decompose, and ferment. Chemical properties are very useful in
identifying substances. However, unlike physical properties, chemical properties can only be observed as the substance is in the
process of being changed into a different substance.
Table 1.7.2 : Contrasting Physical and Chemical Properties
Physical Properties Chemical Properties

Gallium metal melts at 30 oC. Iron metal rusts.

Mercury is a very dense liquid. A green banana turns yellow when it ripens.

Gold is shiny. A dry piece of paper burns.

1.7.2 https://chem.libretexts.org/@go/page/423570
 Example 1.7.1

Which of the following is a chemical property of iron?


a. Iron corrodes in moist air.
b. Density = 7.874 g/cm3
c. Iron is soft when pure.
d. Iron melts at 1808 K.

Solution
"Iron corrodes in moist air" is the only chemical property of iron from the list.

 Exercise 1.7.1A

Which of the following is a physical property of matter?


a. corrosiveness
b. pH (acidity)
c. density
d. flammability

Answer
c

 Exercise 1.7.1B

Which of the following is a chemical property?


a. flammability
b. melting point
c. boiling point
d. density

Answer
a

Summary
A physical property is a characteristic of a substance that can be observed or measured without changing the identity of the
substance. Physical properties include color, density, hardness, and melting and boiling points. A chemical property describes the
ability of a substance to undergo a specific chemical change. To identify a chemical property, we look for a chemical change. A
chemical change always produces one or more types of matter that differ from the matter present before the change. The formation
of rust is a chemical change because rust is a different kind of matter than the iron, oxygen, and water present before the rust
formed.

1.7: Differences in Matter- Physical and Chemical Properties is shared under a CK-12 license and was authored, remixed, and/or curated by
Marisa Alviar-Agnew & Henry Agnew.
3.5: Differences in Matter- Physical and Chemical Properties by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

1.7.3 https://chem.libretexts.org/@go/page/423570
1.8: Changes in Matter - Physical and Chemical Changes
 Learning Objectives
Label a change as chemical or physical.
List evidence that can indicate a chemical change occurred.

Change is happening all around us all of the time. Just as chemists have classified elements and compounds, they have also
classified types of changes. Changes are classified as either physical or chemical changes. Chemists learn a lot about the nature of
matter by studying the changes that matter can undergo. Chemists make a distinction between two different types of changes that
they study—physical changes and chemical changes.

Physical Change
Physical changes are changes in which no bonds are broken or formed. This means that the same types of compounds or elements
that were there at the beginning of the change are there at the end of the change. Because the ending materials are the same as the
beginning materials, the properties (such as color, boiling point, etc.) will also be the same. Physical changes involve moving
molecules around, but not changing them. Some types of physical changes include:
Changes of state (changes from a solid to a liquid or a gas and vice versa).
Separation of a mixture.
Physical deformation (cutting, denting, stretching).
Making solutions (special kinds of mixtures).
As an ice cube melts, its shape changes as it acquires the ability to flow. However, its composition does not change. Melting is an
example of a physical change. A physical change is a change to a sample of matter in which some properties of the material
change, but the identity of the matter does not. When liquid water is heated, it changes to water vapor. However, even though the
physical properties have changed, the molecules are exactly the same as before. We still have each water molecule containing two
hydrogen atoms and one oxygen atom covalently bonded. When you have a jar containing a mixture of pennies and nickels and you
sort the mixture so that you have one pile of pennies and another pile of nickels, you have not altered the identity of the pennies or
the nickels—you've merely separated them into two groups. This would be an example of a physical change. Similarly, if you have
a piece of paper, you don't change it into something other than a piece of paper by ripping it up. What was paper before you started
tearing is still paper when you are done. Again, this is an example of a physical change.

Figure 1.8.1 : Ice melting is a physical change. When liquid water (H O ) freezes into a solid state (ice), it appears changed;
2

however, this change is only physical, as the composition of the constituent molecules is the same: 11.19% hydrogen and 88.81%
oxygen by mass. (Public Domain; Moussa).
Physical changes can further be classified as reversible or irreversible. The melted ice cube may be refrozen, so melting is a
reversible physical change. Physical changes that involve a change of state are all reversible. Other changes of state include
vaporization (liquid to gas), freezing (liquid to solid), and condensation (gas to liquid). Dissolving is also a reversible physical
change. When salt is dissolved into water, the salt is said to have entered the aqueous state. The salt may be regained by boiling off
the water, leaving the salt behind.

1.8.1 https://chem.libretexts.org/@go/page/423571
Chemical Change
Chemical changes occur when bonds are broken and/or formed between molecules or atoms. This means that one substance with a
certain set of properties (such as melting point, color, taste, etc) is turned into a different substance with different properties.
Chemical changes are frequently harder to reverse than physical changes.
One good example of a chemical change is burning a candle. The act of burning paper actually results in the formation of new
chemicals (carbon dioxide and water) from the burning of the wax. Another example of a chemical change is what occurs when
natural gas is burned in your furnace. This time, on the left there is a molecule of methane, CH , and two molecules of oxygen,
4

O ; on the right are two molecules of water, H O , and one molecule of carbon dioxide, CO . In this case, not only has the
2 2 2

appearance changed, but the structure of the molecules has also changed. The new substances do not have the same chemical
properties as the original ones. Therefore, this is a chemical change.

Figure 1.8.2 : Burning of wax to generate water and carbon dioxide is a chemical reaction. (CC-SA-BY-3.0; Andrikkos )
We can't actually see molecules breaking and forming bonds, although that's what defines chemical changes. We have to make
other observations to indicate that a chemical change has happened. Some of the evidence for chemical change will involve the
energy changes that occur in chemical changes, but some evidence involves the fact that new substances with different properties
are formed in a chemical change.
Observations that help to indicate chemical change include:
Temperature changes (either the temperature increases or decreases).
Light given off.
Unexpected color changes (a substance with a different color is made, rather than just mixing the original colors together).
Bubbles are formed (but the substance is not boiling—you made a substance that is a gas at the temperature of the beginning
materials, instead of a liquid).
Different smell or taste (do not taste your chemistry experiments, though!).
A solid forms if two clear liquids are mixed (look for floaties—technically called a precipitate).

 Example 1.8.1

Label each of the following changes as a physical or chemical change. Give evidence to support your answer.
a. Boiling water.
b. A nail rusting.
c. A green solution and colorless solution are mixed. The resulting mixture is a solution with a pale green color.
d. Two colorless solutions are mixed. The resulting mixture has a yellow precipitate.

Solution
a. Physical: boiling and melting are physical changes. When water boils, no bonds are broken or formed. The change could be
written: H O (l) → H O (g)
2 2

b. Chemical: The dark grey nail changes color to form an orange flaky substance (the rust); this must be a chemical change.
Color changes indicate chemical change. The following reaction occurs: Fe + O → Fe O 2 2 3

1.8.2 https://chem.libretexts.org/@go/page/423571
c. Physical: because none of the properties changed, this is a physical change. The green mixture is still green and the
colorless solution is still colorless. They have just been spread together. No color change occurred or other evidence of
chemical change.
d. Chemical: the formation of a precipitate and the color change from colorless to yellow indicate a chemical change.

 Exercise 1.8.1

Label each of the following changes as a physical or chemical change.


a. A mirror is broken.
b. An iron nail corroded in moist air
c. Copper metal is melted.
d. A catalytic converter changes nitrogen dioxide to nitrogen gas and oxygen gas.

Answer a:
physical change

Answer b:
chemical change

Answer c:
physical change

Answer d:
chemical change

Separating Mixtures Through Physical Changes


Homogeneous mixtures (solutions) can be separated into their component substances by physical processes that rely on differences
in some physical property, such as differences in their boiling points. Two of these separation methods are distillation and
crystallization. Distillation makes use of differences in volatility, a measure of how easily a substance is converted to a gas at a
given temperature. A simple distillation apparatus for separating a mixture of substances, at least one of which is a liquid. The most
volatile component boils first and is condensed back to a liquid in the water-cooled condenser, from which it flows into the
receiving flask. If a solution of salt and water is distilled, for example, the more volatile component, pure water, collects in the
receiving flask, while the salt remains in the distillation flask.

1.8.3 https://chem.libretexts.org/@go/page/423571
Figure 1.8.3 : The Distillation of a Solution of Table Salt in Water. The solution of salt in water is heated in the distilling flask until
it boils. The resulting vapor is enriched in the more volatile component (water), which condenses to a liquid in the cold condenser
and is then collected in the receiving flask.
Parts of a distillation setup: Bunsen burner, salt water in distilling flask, condenser with cool water in and warm water out, pure
water in receiving flask
Mixtures of two or more liquids with different boiling points can be separated with a more complex distillation apparatus. One
example is the refining of crude petroleum into a range of useful products: aviation fuel, gasoline, kerosene, diesel fuel, and
lubricating oil (in the approximate order of decreasing volatility). Another example is the distillation of alcoholic spirits such as
brandy or whiskey. This relatively simple procedure caused more than a few headaches for federal authorities in the 1920s during
the era of Prohibition, when illegal stills proliferated in remote regions of the United States.
Another example for using physical properties to separate mixtures is filtration (Figure 1.8.4). Filtration is any mechanical,
physical or biological operation that separates solids from fluids (liquids or gases) by adding a medium through which only the
fluid can pass. The fluid that passes through is called the filtrate. There are many different methods of filtration; all aim to attain the
separation of substances. Separation is achieved by some form of interaction between the substance or objects to be removed and
the filter. The substance that is to pass through the filter must be a fluid, i.e. a liquid or gas. Methods of filtration vary depending on
the location of the targeted material, i.e. whether it is dissolved in the fluid phase or suspended as a solid.

1.8.4 https://chem.libretexts.org/@go/page/423571
Figure 1.8.4 : Filtration for the separation of solids from a hot solution. (CC BY-SA 4.0; Suman6395).

Summary
Chemists make a distinction between two different types of changes that they study—physical changes and chemical changes.
Physical changes are changes that do not alter the identity of a substance.
Chemical changes are changes that occur when one substance is turned into another substance.
Chemical changes are frequently harder to reverse than physical changes. Observations that indicate a chemical change has
occurred include color change, temperature change, light given off, formation of bubbles, formation of a precipitate, etc.

Contributions & Attributions

Boundless (www.boundless.com)

1.8: Changes in Matter - Physical and Chemical Changes is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa
Alviar-Agnew & Henry Agnew.
3.6: Changes in Matter - Physical and Chemical Changes by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

1.8.5 https://chem.libretexts.org/@go/page/423571
1.9: Conservation of Mass - There is No New Matter
It may seem as though burning destroys matter, but the same amount, or mass, of matter still exists after a campfire as before. Look
at Figure 1.9.1 below. It shows that when wood burns, it combines with oxygen and changes not only to ashes, but also to carbon
dioxide and water vapor. The gases float off into the air, leaving behind just the ashes. Suppose you had measured the mass of the
wood before it burned and the mass of the ashes after it burned. Also suppose you had been able to measure the oxygen used by the
fire and the gases produced by the fire. What would you find? The total mass of matter after the fire would be the same as the total
mass of matter before the fire.

Figure 1.9.1 : Burning is a chemical process. The flames are caused as a result of a fuel undergoing combustion (burning). (CC BY-
SA 2.5; Einar Helland Berger for fire and Walter Siegmund for ash).

Law of Conservation of Mass


The law of conservation of mass was created in 1789 by a French chemist, Antoine Lavoisier. The law of conservation of mass
states that matter cannot be created or destroyed in a chemical reaction. For example, when wood burns, the mass of the soot,
ashes, and gases equals the original mass of the charcoal and the oxygen when it first reacted. So the mass of the product equals the
mass of the reactant. A reactant is the chemical reaction of two or more elements to make a new substance, and a product is the
substance that is formed as the result of a chemical reaction (Video 1.9.1). Matter and its corresponding mass may not be able to be
created or destroyed, but can change forms to other substances like liquids, gases, and solids.

Demo - Conservation of Matter

Video 1.9.1 : This is a nice little demonstration showing the Conservation of Mass in action.
If you witness a 300 kg tree burn to the ground, there are only ashes left after the burn, and all of them together weigh 10 kg. It may
make you wonder where the other 290 kg went. The missing 290 kg was released into the atmosphere as smoke, so the only thing
left that you can see is the 10 kg of ash. If you know the law of conservation of mass, then you know that the other 290 kg has to go
somewhere, because it has to equal the mass of the tree before it burnt down.

 Example 1.9.1

If heating 10.0 grams of calcium carbonate (CaCO3) produces 4.4 g of carbon dioxide (CO2) and 5.6 g of calcium oxide (CaO),
show that these observations are in agreement with the law of conservation of mass.

1.9.1 https://chem.libretexts.org/@go/page/423572
Solution
Mass of the reactants = Mass of the products

10.0 g of CaCO = 4.4 g of CO + 5.6 g of CaO


3 2

10.0 g of reactant = 10.0 g of products

Because the mass of the reactant is equal to the mass of the products, the observations are in agreement with the law of
conservation of mass.

 Exercise 1.9.1

Potassium hydroxide (KOH ) readily reacts with carbon dioxide (CO ) to produce potassium carbonate (K CO ) and water (
2 2 3

H O ). How many grams of potassium carbonate are produced if 224.4 g of KOH reacts with 88.0 g of CO ? The reaction
2 2

also produces 36.0 g of water.

Answer
276.4 g of potassium carbonate

The Law is also applicable to both chemical and physical changes. For example, if you have an ice cube that melts into a liquid and
you heat that liquid up, it becomes a gas. It will appear to have disappeared, but is still there.

Summary
Burning and other changes in matter do not destroy matter.
The mass of matter is always the same before and after the changes occur.
The law of conservation of mass states that matter cannot be created or destroyed.

Contributions & Attributions


Binod Shrestha (University of Lorraine)

1.9: Conservation of Mass - There is No New Matter is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa
Alviar-Agnew & Henry Agnew.
3.7: Conservation of Mass - There is No New Matter by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

1.9.2 https://chem.libretexts.org/@go/page/423572
1.10: Compounds Display Constant Composition
When building a house, the starting point is a blueprint of what the house will look like. The plan states how many windows and
what kind, how many doors and what style, how many rooms and what type (bedroom, kitchen, other). The blueprint shows how
the different pieces will go together to make the house. As long as the blueprint is followed and exactly the same items are used,
the house will be identical to its blueprint.

Compounds
A compound is a substance that contains two or more elements chemically combined in a fixed proportion. The elements carbon
and hydrogen combine to form many different compounds. One of the simplest is called methane, in which there are always four
times as many hydrogen particles as carbon particles. Methane is a pure substance because it always has the same composition.
However, it is not an element because it can be broken down into simpler substances—carbon and hydrogen.
Recall that the components of a mixture can be separated from one another by physical means. This is not true for a compound.
Table salt is a compound consisting of equal parts of the elements sodium and chlorine. Salt cannot be separated into its two
elements by filtering, distillation, or any other physical process. Salt and other compounds can only be decomposed into their
elements by a chemical process. A chemical change is a change that produces matter with a different composition. Many
compounds can be decomposed into their elements by heating. When sugar is heated, it decomposes into carbon and water. Water is
still a compound, but one which cannot be broken down into hydrogen and oxygen by heating. Instead, the passage of an electrical
current through water will produce hydrogen and oxygen gases.
The properties of compounds are generally very different than the properties of the elements from which the compound is formed.
Sodium is an extremely reactive soft metal that cannot be exposed to air or water. Chlorine is a deadly gas. The compound sodium
chloride is a white solid which is essential for all living things (see below).

Figure 1.10.1 : (A) Sodium is so reactive that it must be stored under oil. (B) Chlorine is a poisonous yellow-green gas. (C) Salt
crystals, a compound of sodium and chlorine.

Summary
A compound is a substance that contains two or more elements chemically combined in a fixed proportion.
A chemical change is a change that produces matter with a different composition.

1.10: Compounds Display Constant Composition is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-
Agnew & Henry Agnew.
5.2: Compounds Display Constant Composition by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

1.10.1 https://chem.libretexts.org/@go/page/423573
1.11: Chemical Formulas - How to Represent Compounds
 Learning Objectives
Determine the number of different atoms in a formula.
Define chemical formula, molecular formula, and empirical formula.

A chemical formula is an expression that shows the elements in a compound and the relative proportions of those elements. Water
is composed of hydrogen and oxygen in a 2:1 ratio. The chemical formula for water is H O . Sulfuric acid is one of the most widely
2

produced chemicals in the United States and is composed of the elements hydrogen, sulfur, and oxygen. The chemical formula for
sulfuric acid is H SO .
2 4

Certain groups of atoms are bonded together to form what is called a polyatomic ion that acts as a single unit. Polyatomic ions are
discussed in more detail in Section 5.5. Polyatomic ions are enclosed in parenthesis followed by a subscript if more than one of the
same ion exist in a chemical formula. The formula Ca (PO ) represents a compound with the following:
3 4 2

3 Ca atoms + 2 PO43- ions


To count the total number of atoms for formulas with polyatomic ions enclosed in parenthesis, use the subscript as a multiplier for
each atom or number of atoms.

Ca3(PO4)2
3 Ca + 2 x1 P + 2 x 4 O = 3 Ca atoms + 2 P atoms + 8 O atoms
Molecular Formula
A molecular formula is a chemical formula of a molecular compound that shows the kinds and numbers of atoms present in a
molecule of the compound. Ammonia is a compound of nitrogen and hydrogen as shown below:

Figure 1.11.1 : The molecular formula for ammonia. NH3. There is one atom of nitrogen and 3 atoms of hydrogen in a molecule of
ammonia.
Note from the example that there are some standard rules to follow in writing molecular formulas. The arrangements of the
elements depend on the particular structure, which is not of concern at this point. The number of atoms of each kind is indicated by
a subscript following the atom. If there is only one atom, no number is written. If there is more than one atom of a specific kind, the
number is written as a subscript following the atom. We would not write N H for ammonia, because that would mean that there are
3

three nitrogen atoms and one hydrogen atom in the molecule, which is incorrect.

Empirical Formula
An empirical formula is a formula that shows the elements in a compound in their lowest whole-number ratio. Glucose is an
important simple sugar that cells use as their primary source of energy. Its molecular formula is C H O . Since each of the
6 12 6

subscripts is divisible by 6, the empirical formula for glucose is CH O. When chemists analyze an unknown compound, often the
2

first step is to determine its empirical formula.


molecular formula: C H O
6 12 6

empirical formula: CH O 2

There are a great many compounds whose molecular and empirical formulas are the same. If the molecular formula cannot be
simplified into a smaller whole-number ratio, as in the case of H O or P O , then the empirical formula is also the molecular
2 2 5

formula.

1.11.1 https://chem.libretexts.org/@go/page/423574
Summary
A chemical formula is an expression that shows the elements in a compound and the relative proportions of those elements.
If only one atom of a specific type is present, no subscript is used.
For atoms that have two or more of a specific type of atom present, a subscript is written after the symbol for that atom.
Polyatomic ions in chemical formulas are enclosed in parentheses followed by a subscript if more than one of the same type of
polyatomic ion exist.
Molecular formulas do not indicate how the atoms are arranged in the molecule.
The empirical formula tells the lowest whole-number ratio of elements in a compound. The empirical formula does not show
the actual number of atoms.

1.11: Chemical Formulas - How to Represent Compounds is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa
Alviar-Agnew & Henry Agnew.
5.3: Chemical Formulas - How to Represent Compounds by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

1.11.2 https://chem.libretexts.org/@go/page/423574
1.12: Energy
 Learning Objectives
Define heat and work.
Distinguish between kinetic energy and potential energy.
State the law of conservation of matter and energy.

Just like matter, energy is a term that we are all familiar with and use on a daily basis. Before you go on a long hike, you eat an
energy bar; every month, the energy bill is paid; on TV, politicians argue about the energy crisis. But what is energy? If you stop to
think about it, energy is very complicated. When you plug a lamp into an electric socket, you see energy in the form of light, but
when you plug a heating pad into that same socket, you only feel warmth. Without energy, we couldn't turn on lights, we couldn't
brush our teeth, we couldn't make our lunch, and we couldn't travel to school. In fact, without energy, we couldn't even wake up
because our bodies require energy to function. We use energy for every single thing that we do, whether we are awake or asleep.

Ability to Do Work or Produce Heat


When we speak of using energy, we are really referring to transferring energy from one place to another. When you use energy to
throw a ball, you transfer energy from your body to the ball, and this causes the ball to fly through the air. When you use energy to
warm your house, you transfer energy from the furnace to the air in your home, and this causes the temperature in your house to
rise. Although energy is used in many kinds of different situations, all of these uses rely on energy being transferred in one of two
ways. Energy can be transferred as heat or as work.
When scientists speak of heat, they are referring to energy that is transferred from an object with a higher temperature to an object
with a lower temperature, as a result of the temperature difference. Heat will "flow" from the hot object to the cold object until both
end up at the same temperature. When you cook with a metal pot, you witness energy being transferred in the form of heat.
Initially, only the stove element is hot—the pot and the food inside the pot are cold. As a result, heat moves from the hot stove
element to the cold pot. After a while, enough heat is transferred from the stove to the pot, raising the temperature of the pot and all
of its contents (Figure 1.12.1).

Figure 1.12.1 : Energy is transferred as heat from the hot stove element to the cooler pot until the pot and its contents become just
as hot as the element. The energy that is transferred into the pot as heat is then used to cook the food.
Heat is only one way in which energy can be transferred. Energy can also be transferred as work. The scientific definition of work
is force (any push or pull) applied over a distance. When you push an object and cause it to move, you do work, and you transfer
some of your energy to the object. At this point, it's important to warn you of a common misconception. Sometimes we think that
the amount of work done can be measured by the amount of effort put in. This may be true in everyday life, but it is not true in
science. By definition, scientific work requires that force be applied over a distance. It does not matter how hard you push or how
hard you pull. If you have not moved the object, you haven't done any work.
So far, we've talked about the two ways in which energy can be transferred from one place, or object, to another. Energy can be
transferred as heat, and energy can be transferred as work. But the question still remains—what IS energy?

Kinetic Energy
Machines use energy, our bodies use energy, energy comes from the sun, energy comes from volcanoes, energy causes forest fires,
and energy helps us to grow food. With all of these seemingly different types of energy, it's hard to believe that there are really only

1.12.1 https://chem.libretexts.org/@go/page/423575
two different forms of energy: kinetic energy and potential energy. Kinetic energy is energy associated with motion. When an
object is moving, it has kinetic energy. When the object stops moving, it has no kinetic energy. While all moving objects have
kinetic energy, not all moving objects have the same amount of kinetic energy. The amount of kinetic energy possessed by an
object is determined by its mass and its speed. The heavier an object is and the faster it is moving, the more kinetic energy it has.
Kinetic energy is very common, and it's easy to spot examples of it in the world around you. Sometimes we even try to capture
kinetic energy and use it to power things like our home appliances. If you are from California, you might have driven through the
Tehachapi Pass near Mojave or the Montezuma Hills in Solano County and seen the windmills lining the slopes of the mountains
(Figure 1.12.2). These are two of the larger wind farms in North America. As wind rushes along the hills, the kinetic energy of the
moving air particles turns the windmills, trapping the wind's kinetic energy so that people can use it in their houses and offices.

Figure 1.12.2 : A wind farm in Solano County harnesses the kinetic energy of the wind. (CC BY-SA 3.0 Unported; BDS2006 at
Wikipedia)

Potential Energy
Potential energy is stored energy. It is energy that remains available until we choose to use it. Think of a battery in a flashlight. If
left on, the flashlight battery will run out of energy within a couple of hours, and the flashlight will die. If, however, you only use
the flashlight when you need it, and turn it off when you don’t, the battery will last for days or even months. The battery contains a
certain amount of energy, and it will power the flashlight for a certain amount of time, but because the battery stores potential
energy, you can choose to use the energy all at once, or you can save it and only use a small amount at a time.
Any stored energy is potential energy. There are a lot of different ways in which energy can be stored, and this can make potential
energy very difficult to recognize. In general, an object has potential energy because of its position relative to another object. For
example, when a rock is held above the earth, it has potential energy because of its position relative to the ground. This is potential
energy because the energy is stored for as long as the rock is held in the air. Once the rock is dropped, though, the stored energy is
released as kinetic energy as the rock falls.

Chemical Energy
There are other common examples of potential energy. A ball at the top of a hill stores potential energy until it is allowed to roll to
the bottom. When two magnets are held next to one another, they store potential energy too. For some examples of potential energy,
though, it's harder to see how "position" is involved. In chemistry, we are often interested in what is called chemical potential
energy. Chemical potential energy is energy stored in the atoms, molecules, and chemical bonds that make up matter. How does
this depend on position?
As you learned earlier, the world, and all of the chemicals in it are made up of atoms and molecules. These store potential energy
that is dependent on their positions relative to one another. Of course, you can't see atoms and molecules. Nevertheless, scientists
do know a lot about the ways in which atoms and molecules interact, and this allows them to figure out how much potential energy
is stored in a specific quantity (like a cup or a gallon) of a particular chemical. Different chemicals have different amounts of
potential energy because they are made up of different atoms, and those atoms have different positions relative to one another.

Since different chemicals have different amounts of potential energy, scientists will sometimes say that potential energy
depends not only on position, but also on composition. Composition affects potential energy because it determines which

1.12.2 https://chem.libretexts.org/@go/page/423575
molecules and atoms end up next to one another. For example, the total potential energy in a cup of pure water is different than
the total potential energy in a cup of apple juice, because the cup of water and the cup of apple juice are composed of different
amounts of different chemicals.

At this point, you may wonder just how useful chemical potential energy is. If you want to release the potential energy stored in an
object held above the ground, you just drop it. But how do you get potential energy out of chemicals? It's actually not difficult. Use
the fact that different chemicals have different amounts of potential energy. If you start with chemicals that have a lot of potential
energy and allow them to react and form chemicals with less potential energy, all the extra energy that was in the chemicals at the
beginning, but not at the end, is released.

Units of Energy
Energy is measured in one of two common units: the calorie and the joule. The joule (J) is the SI unit of energy. The calorie is
familiar because it is commonly used when referring to the amount of energy contained within food. A calorie (cal) is the quantity
of heat required to raise the temperature of 1 gram of water by 1 C. For example, raising the temperature of 100 g of water from
o

20 C to 22 C would require 100 × 2 = 200 cal.


o o

Calories contained within food are actually kilocalories (kcal). In other words, if a certain snack contains 85 food calories, it
actually contains 85 kcal or 85, 000 cal. In order to make the distinction, the dietary calorie is written with a capital C.

1 kilocalorie = 1 Calorie = 1000 calories

To say that the snack "contains" 85 Calories means that 85 kcal of energy are released when that snack is processed by your body.
Heat changes in chemical reactions are typically measured in joules rather than calories. The conversion between a joule and a
calorie is shown below.

1 J = 0.2390 cal or 1 cal = 4.184 J

We can calculate the amount of heat released in kilojoules when a 400 Calorie hamburger is digested.
4.184 kJ 3
400 Cal = 400 kcal × = 1.67 × 10 kJ
1 kcal

Summary
Any time we use energy, we transfer energy from one object to another.
Energy can be transferred in one of two ways: as heat, or as work.
Heat is the term given to energy that is transferred from a hot object to a cooler object due to the difference in their
temperatures.
Work is the term given to energy that is transferred as a result of a force applied over a distance.
Energy comes in two fundamentally different forms: kinetic energy and potential energy.
Kinetic energy is the energy of motion.
Potential energy is stored energy that depends on the position of an object relative to another object.
Chemical potential energy is a special type of potential energy that depends on the positions of different atoms and molecules
relative to one another.
Chemical potential energy can also be thought of according to its dependence on chemical composition.
Energy can be converted from one form to another. The total amount of mass and energy in the universe is conserved.

Contributions & Attributions

Wikibooks

1.12: Energy is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew & Henry Agnew.
3.8: Energy by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source: https://www.ck12.org/c/chemistry/.

1.12.3 https://chem.libretexts.org/@go/page/423575
1.13: Energy and Chemical and Physical Change
 Learning Objectives
Define endothermic and exothermic reactions.
Describe how heat is transferred in endothermic and exothermic reactions.
Determine whether a reaction is endothermic or exothermic through observations, temperature changes, or an energy
diagram.

So far, we have talked about how energy exists as either kinetic energy or potential energy and how energy can be transferred as
either heat or work. While it's important to understand the difference between kinetic energy and potential energy and the
difference between heat and work, the truth is, energy is constantly changing. Kinetic energy is constantly being turned into
potential energy, and potential energy is constantly being turned into kinetic energy. Likewise, energy that is transferred as work
might later end up transferred as heat, while energy that is transferred as heat might later end up being used to do work.
Even though energy can change form, it must still follow one fundamental law: Energy cannot be created or destroyed, it can only
be changed from one form to another. This law is known as the Law of Conservation of Energy. In a lot of ways, energy is like
money. You can exchange quarters for dollar bills and dollar bills for quarters, but no matter how often you convert between the
two, you will not end up with any more or any less money than you started with. Similarly, you can transfer (or spend) money
using cash, or transfer money using a credit card; but you still spend the same amount of money, and the store still makes the same
amount of money.

A campfire is an example of basic thermochemistry. The reaction is initiated by the application of heat from a match. The
reaction converting wood to carbon dioxide and water (among other things) continues, releasing heat energy in the process.
This heat energy can then be used to cook food, roast marshmallows, or just keep warm when it's cold outside.

An image of a campfire with colored flames, made by the burning of a garden hose in a copper pipe. (CC SA-BY 3.0; Jared)

Exothermic and Endothermic Processes


When physical or chemical changes occur, they are generally accompanied by a transfer of energy. The law of conservation of
energy states that in any physical or chemical process, energy is neither created nor destroyed. In other words, the entire energy in
the universe is conserved. In order to better understand the energy changes taking place during a reaction, we need to define two
parts of the universe: the system and the surroundings. The system is the specific portion of matter in a given space that is being
studied during an experiment or an observation. The surroundings are everything in the universe that is not part of the system. In
practical terms for a laboratory chemist, the system is the particular chemicals being reacted, while the surroundings are the
immediate vicinity within the room. During most processes, energy is exchanged between the system and the surroundings. If the
system loses a certain amount of energy, that same amount of energy is gained by the surroundings. If the system gains a certain
amount of energy, that energy is supplied by the surroundings.
A chemical reaction or physical change is endothermic if heat is absorbed by the system from the surroundings. In the course of an
endothermic process, the system gains heat from the surroundings and so the temperature of the surroundings decreases. The
quantity of heat for a process is represented by the letter q. The sign of q for an endothermic process is positive because the system

1.13.1 https://chem.libretexts.org/@go/page/423576
is gaining heat. A chemical reaction or physical change is exothermic if heat is released by the system into the surroundings.
Because the surroundings are gaining heat from the system, the temperature of the surroundings increases. The sign of q for an
exothermic process is negative because the system is losing heat.

Figure 1.13.1 : (A) Endothermic reaction. (B) Exothermic reaction.


Endothermic reaction: surroundings get cooler and delta H is greater than 0, Exothermic reaction: surroundings get warmer and
delta H is less than 0
During phase changes, energy changes are usually involved. For example, when solid dry ice vaporizes (physical change), carbon
dioxide molecules absorb energy. When liquid water becomes ice, energy is released. Remember that all chemical reactions involve
a change in the bonds of the reactants. The bonds in the reactants are broken and the bonds of the products are formed. Chemical
bonds have potential energy or "stored energy". Because we are changing the bonding, this means we are also changing how much
of this "stored energy" there is in a reaction.
Energy changes are frequently shown by drawing an energy diagram. Energy diagrams show the stored/hidden energy of the
reactants and products as well as the activation energy. If, on an energy diagram, the products have more stored energy than the
reactants started with, the reaction is endothermic. You had to give the reaction energy. If, on the energy diagram, the products have
less stored energy than the reactants started with, the reaction is exothermic.

 Example 1.13.1

Label each of the following processes as endothermic or exothermic.


a. water boiling
b. gasoline burning
c. ice forming on a pond

Solution
a. Endothermic—you must put a pan of water on the stove and give it heat in order to get water to boil. Because you are
adding heat/energy, the reaction is endothermic.
b. Exothermic—when you burn something, it feels hot to you because it is giving off heat into the surroundings.
c. Exothermic—think of ice forming in your freezer instead. You put water into the freezer, which takes heat out of the water,
to get it to freeze. Because heat is being pulled out of the water, it is exothermic. Heat is leaving.

 Exercise 1.13.1

Label each of the following processes as endothermic or exothermic.


a. water vapor condensing
b. gold melting

Answer (a)
exothermic

Answer (b)
endothermic

1.13.2 https://chem.libretexts.org/@go/page/423576
Summary
Phase changes involve changes in energy. All chemical reactions involve changes in energy. This may be a change in heat,
electricity, light, or other forms of energy. Reactions that absorb energy are endothermic. Reactions that release energy are
exothermic.

1.13: Energy and Chemical and Physical Change is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-
Agnew & Henry Agnew.
3.9: Energy and Chemical and Physical Change by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

1.13.3 https://chem.libretexts.org/@go/page/423576
1.14: TAKING MEASUREMENTS
 LEARNING OBJECTIVES
Express quantities properly, using a number and a unit.

A coffee maker’s instructions tell you to fill the coffee pot with 4 cups of water and to use 3 scoops of coffee. When you follow these
instructions, you are measuring. When you visit a doctor’s office, a nurse checks your temperature, height, weight, and perhaps blood
pressure (Figure 1.14.1 ); the nurse is also measuring.

Figure 1.14.1: Measuring Blood Pressure. A nurse or a doctor measuring a patient’s blood pressure is taking a measurement. (GFDL; Pia
von Lützau).
Chemists measure the properties of matter and express these measurements as quantities. A quantity is an amount of something and consists
of a number and a unit. The number tells us how many (or how much), and the unit tells us what the scale of measurement is. For example,
when a distance is reported as “5 kilometers,” we know that the quantity has been expressed in units of kilometers and that the number of
kilometers is 5. If you ask a friend how far they walk from home to school, and the friend answers “12” without specifying a unit, you do
not know whether your friend walks 12 kilometers, 12 miles, 12 furlongs, or 12 yards. Both a number and a unit must be included to express
a quantity properly.
To understand chemistry, we need a clear understanding of the units chemists work with and the rules they follow for expressing numbers.
The next two sections examine the rules for expressing numbers.

 EXAMPLE 1.14.1

Identify the number and the unit in each quantity.


a. one dozen eggs
b. 2.54 centimeters
c. a box of pencils
d. 88 meters per second

SOLUTION
a. The number is one, and the unit is a dozen eggs.
b. The number is 2.54, and the unit is centimeter.
c. The number 1 is implied because the quantity is only a box. The unit is box of pencils.
d. The number is 88, and the unit is meters per second. Note that in this case the unit is actually a combination of two units: meters and
seconds.

KEY TAKE AWAY


Identify a quantity properly with a number and a unit.

1.14: Taking Measurements is shared under a CC BY-NC-SA 3.0 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew & Henry
Agnew.
2.1: Taking Measurements by Henry Agnew, Marisa Alviar-Agnew is licensed CC BY-NC-SA 3.0.

1.14.1 https://chem.libretexts.org/@go/page/423577
1.15: Scientific Notation - Writing Large and Small Numbers
 Learning Objectives
Express a large number or a small number in scientific notation.
Carry out arithmetical operations and express the final answer in scientific notation

Chemists often work with numbers that are exceedingly large or small. For example, entering the mass in grams of a hydrogen
atom into a calculator would require a display with at least 24 decimal places. A system called scientific notation avoids much of
the tedium and awkwardness of manipulating numbers with large or small magnitudes. In scientific notation, these numbers are
expressed in the form
n
N × 10

where N is greater than or equal to 1 and less than 10 (1 ≤ N < 10), and n is a positive or negative integer (100 = 1). The number 10
is called the base because it is this number that is raised to the power n . Although a base number may have values other than 10,
the base number in scientific notation is always 10.
A simple way to convert numbers to scientific notation is to move the decimal point as many places to the left or right as needed to
give a number from 1 to 10 (N). The magnitude of n is then determined as follows:
If the decimal point is moved to the left n places, n is positive.
If the decimal point is moved to the right n places, n is negative.
Another way to remember this is to recognize that as the number N decreases in magnitude, the exponent increases and vice versa.
The application of this rule is illustrated in Example 1.15.1.

 Example 1.15.1: Expressing Numbers in Scientific Notation

Convert each number to scientific notation.


a. 637.8
b. 0.0479
c. 7.86
d. 12,378
e. 0.00032
f. 61.06700
g. 2002.080
h. 0.01020

Solution
Solutions to Example 2.2.1
Explanation Answer

To convert 637.8 to a number from 1 to 10,


we move the decimal point two places to the
a left: 637.8 6.378 × 10
2

Because the decimal point was moved two


places to the left, n = 2.

To convert 0.0479 to a number from 1 to 10,


we move the decimal point two places to the
b right: 0.0479 4.79 × 10
−2

Because the decimal point was moved two


places to the right, n = −2.

This is usually expressed simply as 7.86.


c 7.86 × 10
0

(Recall that 100 = 1.)

1.15.1 https://chem.libretexts.org/@go/page/423578
Explanation Answer

Because the decimal point was moved four


d 1.2378 × 10
4

places to the left, n = 4.

Because the decimal point was moved four


e 3.2 × 10
−4

places to the right, n = −4.

Because the decimal point was moved one


f 6.106700 × 10
1

place to the left, n = 1.

Because the decimal point was moved three


g 2.002080 × 10
3

places to the left, n = 3.

Because the decimal point was moved two


h 1.020 × 10
−2

places to the right, n = -2.

Addition and Subtraction


Before numbers expressed in scientific notation can be added or subtracted, they must be converted to a form in which all the
exponents have the same value. The appropriate operation is then carried out on the values of N. Example 1.15.2 illustrates how to
do this.

 Example 1.15.2: Expressing Sums and Differences in Scientific Notation

Carry out the appropriate operation and then express the answer in scientific notation.
a. (1.36 × 10 ) + (4.73 × 10 )
2 3

b. (6.923 × 10 ) − (8.756 × 10
−3 −4
)

Solution
Solutions to Example 2.2.2.
Explanation Answer

Both exponents must have the same value, so


these numbers are converted to either
2 2 2 2
(1.36 × 10 ) + (47.3 × 10 ) = (1.36 + 47.3) × 10 = 48.66 × 10

or
3 3 3 3
(0.136 × 10 ) + (4.73 × 10 ) = (0.136 + 4.73) × 10 ) = 4.87 × 10

a . 4.87 × 10
3

Choosing either alternative gives the same


answer, reported to two decimal places.
In converting 48.66 × 102 to scientific
notation, n has become more positive by 1
because the value of N has decreased.

Converting the exponents to the same value


gives either
−3 −3 −3
(6.923 × 10 ) − (0.8756 × 10 ) = (6.923 − 0.8756) × 10

or
b (69.23 × 10
−4
) − (8.756 × 10
−4
) = (69.23 −6.047
8.756)
××10
10
−3−4
= 60.474 × 10
−4

.
In converting 60.474 × 10-4 to scientific
notation, n has become more positive by 1
because the value of N has decreased.

Multiplication and Division


When multiplying numbers expressed in scientific notation, we multiply the values of N and add together the values of n .
Conversely, when dividing, we divide N in the dividend (the number being divided) by N in the divisor (the number by which we

1.15.2 https://chem.libretexts.org/@go/page/423578
are dividing) and then subtract n in the divisor from n in the dividend. In contrast to addition and subtraction, the exponents do not
have to be the same in multiplication and division. Examples of problems involving multiplication and division are shown in
Example 1.15.3.

 Example 1.15.3: Expressing Products and Quotients in Scientific Notation

Perform the appropriate operation and express your answer in scientific notation.
a. (6.022 × 10
23
)(6.42 × 10
−2
)
−24
1.67 × 10
b. −28
9.12 × 10
−34
(6.63 × 10 )(6.0 × 10)
c. −2
8.52 × 10

Solution
Solution to Example 2.2.3
Explanation Answer

In multiplication, we add the exponents:


23 −2 [23+(−2)] 21
(6.022 × 10 )(6.42 × 10 ) = (6.022)(6.42) × 10 = 38.7 × 10

a
In converting 38.7 × 10 to scientific
21

notation, n has become more positive by 1


because the value of N has decreased.

3.87 × 10
22
b
In division, we subtract the exponents:
−24
1.67 × 10 1.67 [−24−(−28)] 4
= × 10 = 0.183 × 10
−28
9.12 × 10 9.12

In converting 0.183 × 10 to scientific notation, n has become more negative by 1 because the value of N has increased.
4

1.83 × 10
3
c
This problem has both multiplication and division:
−34
(6.63 × 10 )(6.0 × 10) 39.78
[−34+1−(−2)]
= × 10
−2
(8.52 × 10 ) 8.52

−31
4.7 × 10

1.15: Scientific Notation - Writing Large and Small Numbers is shared under a CK-12 license and was authored, remixed, and/or curated by
Marisa Alviar-Agnew & Henry Agnew.
2.2: Scientific Notation - Writing Large and Small Numbers by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

1.15.3 https://chem.libretexts.org/@go/page/423578
1.16: Significant Figures - Writing Numbers to Reflect Precision
 Learning Objectives
Identify the number of significant figures in a reported value.

The significant figures in a measurement consist of all the certain digits in that measurement plus one uncertain or estimated digit.
In the ruler illustration below, the bottom ruler gave a length with 2 significant figures, while the top ruler gave a length with 3
significant figures. In a correctly reported measurement, the final digit is significant but not certain. Insignificant digits are not
reported. With either ruler, it would not be possible to report the length at 2.553 cm as there is no possible way that the thousandths
digit could be estimated. The 3 is not significant and would not be reported.

Figure 1.16.1 : Measurement with two different rulers.


Ruler A's measurement can be rounded to 2.55, with 2 certain digits, while Ruler B's measurement of 2.5 has 1 certain digit

Measurement Uncertainty
Some error or uncertainty always exists in any measurement. The amount of uncertainty depends both upon the skill of the
measurer and upon the quality of the measuring tool. While some balances are capable of measuring masses only to the nearest
0.1 g, other highly sensitive balances are capable of measuring to the nearest 0.001 g or even better. Many measuring tools such as

rulers and graduated cylinders have small lines which need to be carefully read in order to make a measurement. Figure 1.16.1
shows two rulers making the same measurement of an object (indicated by the blue arrow).
With either ruler, it is clear that the length of the object is between 2 and 3 cm. The bottom ruler contains no millimeter markings.
With that ruler, the tenths digit can be estimated and the length may be reported as 2.5 cm. However, another person may judge that
the measurement is 2.4 cm or perhaps 2.6 cm. While the 2 is known for certain, the value of the tenths digit is uncertain.
The top ruler contains marks for tenths of a centimeter (millimeters). Now the same object may be measured as 2.55 cm. The
measurer is capable of estimating the hundredths digit because he can be certain that the tenths digit is a 5. Again, another measurer
may report the length to be 2.54 cm or 2.56 cm. In this case, there are two certain digits (the 2 and the 5), with the hundredths digit
being uncertain. Clearly, the top ruler is a superior ruler for measuring lengths as precisely as possible.

 Example 1.16.1: Reporting Measurements to the Proper Number of Significant Figures

Use each diagram to report a measurement to the proper number of significant figures.

a.
b.

1.16.1 https://chem.libretexts.org/@go/page/423579
Ruler measuring a rectangle in units of centimeters, with the rectangle's edge between 1.2 and 1.3 cm marks

Solutions
Solutions to Example 2.3.1
Explanation Answer

The arrow is between 4.0 and 5.0, so the


measurement is at least 4.0. The arrow is
between the third and fourth small tick
marks, so it’s at least 0.3. We will have to
estimate the last place. It looks like about
a. one-third of the way across the space, so let 4.33 psi
us estimate the hundredths place as 3. The
symbol psi stands for “pounds per square
inch” and is a unit of pressure, like air in a
tire. The measurement is reported to three
significant figures.

The rectangle is at least 1.0 cm wide but


certainly not 2.0 cm wide, so the first
significant digit is 1. The rectangle’s width is
past the second tick mark but not the third; if
each tick mark represents 0.1, then the
rectangle is at least 0.2 in the next significant
b. digit. We have to estimate the next place 1.25 cm
because there are no markings to guide us. It
appears to be about halfway between 0.2 and
0.3, so we will estimate the next place to be a
5. Thus, the measured width of the rectangle
is 1.25 cm. The measurement is reported to
three significant figures.

 Exercise 1.16.1

What would be the reported width of this rectangle?

Answer
1.25 cm

When you look at a reported measurement, it is necessary to be able to count the number of significant figures. The table below
details the rules for determining the number of significant figures in a reported measurement. For the examples in the table, assume

1.16.2 https://chem.libretexts.org/@go/page/423579
that the quantities are correctly reported values of a measured quantity.
Table 1.16.1 : Significant Figure Rules
Rule Examples

237 has three significant figures.


1. All nonzero digits in a measurement are significant.
1.897 has four significant figures.

2. Zeros that appear between other nonzero digits (middle zeros) are 39,004 has five significant figures.
always significant. 5.02 has three significant figures.

3. Zeros that appear in front of all of the nonzero digits are called 0.008 has one significant figure.
leading zeros. Leading zeros are never significant. 0.000416 has three significant figures.

4. Zeros that appear after all nonzero digits are called trailing zeros. A 1400 is ambiguous.
number with trailing zeros that lacks a decimal point may or may not be 1.4 × 10 has two significant figures.
3

significant. Use scientific notation to indicate the appropriate 1.40 × 10 three significant figures.
3

number of significant figures. 1.400 × 10 has four significant figures.


3

5. Trailing zeros in a number with a decimal point are significant. This 620.0 has four significant figures.
is true whether the zeros occur before or after the decimal point. 19.000 has five significant figures.

Exact Numbers
Integers obtained either by counting objects or from definitions are exact numbers, which are considered to have infinitely many
significant figures. If we have counted four objects, for example, then the number 4 has an infinite number of significant figures
(i.e., it represents 4.000…). Similarly, 1 foot (ft) is defined to contain 12 inches (in), so the number 12 in the following equation has
infinitely many significant figures:

 Example 1.16.2

Give the number of significant figures in each. Identify the rule for each.
a. 5.87
b. 0.031
c. 52.90
d. 00.2001
e. 500
f. 6 atoms

Solution
Solution to Example 2.3.2
Explanation Answer

a All three numbers are significant (rule 1). 5.87, three significant figures

The leading zeros are not significant (rule 3).


b 0.031, two significant figures
The 3 and the 1 are significant (rule 1).

The 5, the 2 and the 9 are significant (rule 1).


c 52.90, four significant figures
The trailing zero is also significant (rule 5).

The leading zeros are not significant (rule 3).


d The 2 and the 1 are significant (rule 1) and 00.2001, four significant figures
the middle zeros are also significant (rule 2).

The number is ambiguous. It could have one,


e 500, ambiguous
two or three significant figures.

The 6 is a counting number. A counting


f 6, infinite
number is an exact number.

1.16.3 https://chem.libretexts.org/@go/page/423579
 Exercise 1.16.2

Give the number of significant figures in each.


a. 36.7 m
b. 0.006606 s
c. 2,002 kg
d. 306,490,000 people
e. 3,800 g

Answer a
three significant figures

Answer b
four significant figures

Answer c
four significant figures

Answer d
infinite (exact number)

Answer e
Ambiguous, could be two, three or four significant figures.

Accuracy and Precision


Measurements may be accurate, meaning that the measured value is the same as the true value; they may be precise, meaning that
multiple measurements give nearly identical values (i.e., reproducible results); they may be both accurate and precise; or they may
be neither accurate nor precise. The goal of scientists is to obtain measured values that are both accurate and precise. The video
below demonstrates the concepts of accuracy and precision.

What's the difference between accurac…


accurac…

Video 1.16.1 : Difference between precision and accuracy.

 Example 1.16.3
The following archery targets show marks that represent the results of four sets of measurements.

1.16.4 https://chem.libretexts.org/@go/page/423579
Which target shows
a. a precise, but inaccurate set of measurements?
b. a set of measurements that is both precise and accurate?
c. a set of measurements that is neither precise nor accurate?

Solution
a. Set a is precise, but inaccurate.
b. Set c is both precise and accurate.
c. Set d is neither precise nor accurate.

Summary
Uncertainty exists in all measurements. The degree of uncertainty is affected in part by the quality of the measuring tool.
Significant figures give an indication of the certainty of a measurement. Rules allow decisions to be made about how many digits to
use in any given situation.

1.16: Significant Figures - Writing Numbers to Reflect Precision is shared under a CK-12 license and was authored, remixed, and/or curated by
Marisa Alviar-Agnew, Henry Agnew, Sridhar Budhi, & Sridhar Budhi.
2.3: Significant Figures - Writing Numbers to Reflect Precision by Henry Agnew, Marisa Alviar-Agnew, Sridhar Budhi is licensed CK-12.
Original source: https://www.ck12.org/c/chemistry/.

1.16.5 https://chem.libretexts.org/@go/page/423579
1.17: Significant Figures in Calculations
 Learning Objectives
Use significant figures correctly in arithmetical operations.

Rounding
Before dealing with the specifics of the rules for determining the significant figures in a calculated result, we need to be able to
round numbers correctly. To round a number, first decide how many significant figures the number should have. Once you know
that, round to that many digits, starting from the left. If the number immediately to the right of the last significant digit is less than
5, it is dropped and the value of the last significant digit remains the same. If the number immediately to the right of the last
significant digit is greater than or equal to 5, the last significant digit is increased by 1.
Consider the measurement 207.518 m. Right now, the measurement contains six significant figures. How would we successively
round it to fewer and fewer significant figures? Follow the process as outlined in Table 1.17.1.
Table 1.17.1 : Rounding examples
Number of Significant Figures Rounded Value Reasoning

6 207.518 All digits are significant

5 207.52 8 rounds the 1 up to 2

4 207.5 2 is dropped

3 208 5 rounds the 7 up to 8

2 210 8 is replaced by a 0 and rounds the 0 up to 1

1 200 1 is replaced by a 0

Notice that the more rounding that is done, the less reliable the figure is. An approximate value may be sufficient for some
purposes, but scientific work requires a much higher level of detail.
It is important to be aware of significant figures when you are mathematically manipulating numbers. For example, dividing 125 by
307 on a calculator gives 0.4071661238… to an infinite number of digits. But do the digits in this answer have any practical
meaning, especially when you are starting with numbers that have only three significant figures each? When performing
mathematical operations, there are two rules for limiting the number of significant figures in an answer—one rule is for addition
and subtraction, and one rule is for multiplication and division.

In operations involving significant figures, the answer is reported in such a way that it
reflects the reliability of the least precise operation. An answer is no more precise than
the least precise number used to get the answer.

Multiplication and Division


For multiplication or division, the rule is to count the number of significant figures in each number being multiplied or divided and
then limit the significant figures in the answer to the lowest count. An example is as follows:

The final answer, limited to four significant figures, is 4,094. The first digit dropped is 1, so we do not round up.
Scientific notation provides a way of communicating significant figures without ambiguity. You simply include all the significant
figures in the leading number. For example, the number 450 has two significant figures and would be written in scientific notation
as 4.5 × 102, whereas 450.0 has four significant figures and would be written as 4.500 × 102. In scientific notation, all significant
figures are listed explicitly.

1.17.1 https://chem.libretexts.org/@go/page/423580
 Example 1.17.1

Write the answer for each expression using scientific notation with the appropriate number of significant figures.
a. 23.096 × 90.300
b. 125 × 9.000

Solution
a
Table with two columns and 1 row. The first column on the left is labeled, Explanation, and underneath in the row is an explanation. The
second column is labeled, Answer, and underneath in the row is an answer.
Explanation Answer

The calculator answer is 2,085.5688, but we need to round it to five


significant figures. Because the first digit to be dropped (in the tenths 2.0856 × 10
3

place) is greater than 5, we round up to 2,085.6.

b
Table with two columns and 1 row. The first column on the left is labeled, Explanation, and underneath in the row is an explanation. The
second column is labeled, Answer, and underneath in the row is an answer.
Explanation Answer

The calculator gives 1,125 as the answer, but we limit it to three 3


1.13 × 10
significant figures.

Addition and Subtraction


How are significant figures handled in calculations? It depends on what type of calculation is being performed. If the calculation is
an addition or a subtraction, the rule is as follows: limit the reported answer to the rightmost column that all numbers have
significant figures in common. For example, if you were to add 1.2 and 4.71, we note that the first number stops its significant
figures in the tenths column, while the second number stops its significant figures in the hundredths column. We therefore limit our
answer to the tenths column.

We drop the last digit—the 1—because it is not significant to the final answer.
The dropping of positions in sums and differences brings up the topic of rounding. Although there are several conventions, in this
text we will adopt the following rule: the final answer should be rounded up if the first dropped digit is 5 or greater, and rounded
down if the first dropped digit is less than 5.

 Example 1.17.2
a. 13.77 + 908.226
b. 1,027 + 611 + 363.06

1.17.2 https://chem.libretexts.org/@go/page/423580
Solution
a
Table with two columns and 1 row. The first column on the left is labeled, Explanation, and underneath in the row is an explanation. The
second column is labeled, Answer, and underneath in the row is an answer.
Explanation Answer

The calculator answer is 921.996, but because 13.77 has its farthest-
right significant figure in the hundredths place, we need to round the
final answer to the hundredths position. Because the first digit to be 922.00 = 9.2200 × 10
2

dropped (in the thousandths place) is greater than 5, we round up to


922.00

b
Table with two columns and 1 row. The first column on the left is labeled, Explanation, and underneath in the row is an explanation. The
second column is labeled, Answer, and underneath in the row is an answer.
Explanation Answer

The calculator gives 2,001.06 as the answer, but because 611 and
1027 has its farthest-right significant figure in the ones place, the 2, 001.06 = 2.001 × 10
3

final answer must be limited to the ones position.

 Exercise 1.17.2

Write the answer for each expression using scientific notation with the appropriate number of significant figures.
a. 217 ÷ 903
b. 13.77 + 908.226 + 515
c. 255.0 − 99
d. 0.00666 × 321

Answer a:
−1
0.240 = 2.40 × 10

Answer b:
3
1, 437 = 1.437 × 10

Answer c:
2
156 = 1.56 × 10

Answer d:
0
2.14 = 2.14 × 10

Remember that calculators do not understand significant figures. You are the one who must apply the rules of significant
figures to a result from your calculator.

Calculations Involving Multiplication/Division and Addition/Subtraction


In practice, chemists generally work with a calculator and carry all digits forward through subsequent calculations. When working
on paper, however, we often want to minimize the number of digits we have to write out. Because successive rounding can
compound inaccuracies, intermediate rounding needs to be handled correctly. When working on paper, always round an
intermediate result so as to retain at least one more digit than can be justified and carry this number into the next step in the
calculation. The final answer is then rounded to the correct number of significant figures at the very end.

1.17.3 https://chem.libretexts.org/@go/page/423580
Signi cant gures in mixed operations

Video 1.17.1 : Significant figures in mixed operations (https://www.youtube.com/watch?v=yBntMndXQWA).

Signi cant Figures in Combined Operati…


Operati…

Video 1.17.2 : https://www.youtube.com/watch?v=__csP0NtlGI


In the worked examples in this text, we will often show the results of intermediate steps in a calculation. In doing so, we will show
the results to only the correct number of significant figures allowed for that step, in effect treating each step as a separate
calculation. This procedure is intended to reinforce the rules for determining the number of significant figures, but in some cases it
may give a final answer that differs in the last digit from that obtained using a calculator, where all digits are carried through to the
last step.

 Example 1.17.3
a. 2(1.008 g) + 15.99 g
b. 137.3 s + 2(35.45 s)
118.7g
c. 2
− 35.5g

Solution
a.
Table with two columns and 1 row. The first column on the left is labeled, Explanation, and underneath in the row is an explanation for
multiplication first. The second column is labeled, Answer, and underneath in the row is an answer.
Explanation Answer

1.17.4 https://chem.libretexts.org/@go/page/423580
2(1.008 g) + 15.99 g =
Perform multiplication first.
2 (1.008 g 4 sig figs) = 2.016 g 4 sig figs
The number with the least number of significant figures is 1.008 g;
the number 2 is an exact number and therefore has an infinite
number of significant figures.
Then, perform the addition. 18.01 g (rounding up)
2.016 g thousandths place + 15.99 g hundredths place (least precise)
= 18.006 g
Round the final answer.
Round the final answer to the hundredths place since 15.99 has its
farthest right significant figure in the hundredths place (least
precise).

b.
Table with two columns and 1 row. The first column on the left is labeled, Explanation, and underneath in the row is an explanation for
multiplication first. The second column is labeled, Answer, and underneath in the row is an answer.
Explanation Answer

137.3 s + 2(35.45 s) =
Perform multiplication first.
2(35.45 s 4 sig figs) = 70.90 s 4 sig figs
The number with the least number of significant figures is 35.45; the
number 2 is an exact number and therefore has an infinite number of
significant figures. 208.2 s
Then, perform the addition.
137.3 s tenths place (least precise) + 70.90 s hundredths place =
208.20 s
Round the final answer.
Round the final answer to the tenths place based on 137.3 s.

c.
Table with two columns and 1 row. The first column on the left is labeled, Explanation, and underneath in the row is an explanation for
division first. The second column is labeled, Answer, and underneath in the row is an answer.
Explanation Answer
118.7g

2
− 35.5g =
Perform division first.
118.7g

2
4 sig figs = 59.35 g 4 sig figs
The number with the least number of significant figures is 118.7 g;
the number 2 is an exact number and therefore has an infinite
number of significant figures. 23.9 g (rounding up)
Perform subtraction next.
59.35 g hundredths place − 35.5 g tenths place (least precise) =
23.85 g
Round the final answer.
Round the final answer to the tenths place based on 35.5 g.

1.17.5 https://chem.libretexts.org/@go/page/423580
 Exercise 1.17.3

Complete the calculations and report your answers using the correct number of significant figures.
a. 5(1.008s) - 10.66 s
b. 99.0 cm+ 2(5.56 cm)

Answer a
-5.62 s
Answer b
110.2 cm

Summary
Rounding
If the number to be dropped is greater than or equal to 5, increase the number to its left by 1 (e.g. 2.9699 rounded to three
significant figures is 2.97).
If the number to be dropped is less than 5, there is no change (e.g. 4.00443 rounded to four significant figures is 4.004).
The rule in multiplication and division is that the final answer should have the same number of significant figures as there are in
the number with the fewest significant figures.
The rule in addition and subtraction is that the answer is given the same number of decimal places as the term with the fewest
decimal places.

1.17: Significant Figures in Calculations is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew &
Henry Agnew.
2.4: Significant Figures in Calculations by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

1.17.6 https://chem.libretexts.org/@go/page/423580
1.18: The Basic Units of Measurement
 Learning Objectives
State the different measurement systems used in chemistry.
Describe how prefixes are used in the metric system and identify how the prefixes milli-, centi-, and kilo- compare to the
base unit.

How long is a yard? It depends on whom you ask and when you asked the question. Today we have a standard definition of the
yard, which you can see marked on every football field. If you move the ball ten yards, you get a first down and it does not matter
whether you are playing in Los Angeles, Dallas, or Green Bay. But at one time that yard was arbitrarily defined as the distance
from the tip of the king's nose to the end of his outstretched hand. Of course, the problem there is simple: new king, new distance
(and then you have to re-mark all of those football fields).

Figure 1.18.1 : Meter standard (left) and Kilogram standard (right).

SI Base Units
All measurements depend on the use of units that are well known and understood. The English system of measurement units
(inches, feet, ounces, etc.) are not used in science because of the difficulty in converting from one unit to another. The metric
system is used because all metric units are based on multiples of 10, making conversions very simple. The metric system was
originally established in France in 1795. The International System of Units is a system of measurement based on the metric
system. The acronym SI is commonly used to refer to this system and stands for the French term, Le Système International
d'Unités. The SI was adopted by international agreement in 1960 and is composed of seven base units in Table 1.18.1.
Table 1.18.1 : SI Base Units of Measurement
Quantity SI Base Unit Symbol

Length meter m

Mass kilogram kg

Temperature kelvin K

Time second s

Amount of a Substance mole mol

Electric Current ampere A

Luminous Intensity candela cd

The first units are frequently encountered in chemistry. All other measurement quantities, such as volume, force, and energy, can be
derived from these seven base units.

 Unfortunately, the Metric System is Not Ubiquitous


The map below shows the adoption of the SI units in countries around the world. The United States has legally adopted the
metric system for measurements, but does not use it in everyday practice. Great Britain and much of Canada use a combination
of metric and imperial units.

1.18.1 https://chem.libretexts.org/@go/page/423581
Figure 1.18.1 : Areas of world using metric system (in green). Only a few countries are slow or resistant to adoption including
the United States.

Prefix Multipliers
Conversions between metric system units are straightforward because the system is based on powers of ten. For example, meters,
centimeters, and millimeters are all metric units of length. There are 10 millimeters in 1 centimeter and 100 centimeters in 1 meter.
Metric prefixes are used to distinguish between units of different size. These prefixes all derive from either Latin or Greek terms.
For example, mega comes from the Greek word μεγας , meaning "great". Table 1.18.2 lists the most common metric prefixes and
their relationship to the central unit that has no prefix. Length is used as an example to demonstrate the relative size of each
prefixed unit.
Table 1.18.2 : SI Prefixes
Prefix Unit Abbreviation Meaning Example

giga G 1,000,000,000 1 gigameter (Gm) = 10 9


m

mega M 1,000,000 1 megameter (Mm) = 10 6


m

kilo k 1,000 1 kilometer (km) = 1, 000 m

hecto h 100 1 hectometer (hm) = 100 m

deka da 10 1 dekameter (dam) = 10 m

1 1 meter (m)

deci d 1/10 1 decimeter (dm) = 0.1 m

centi c 1/100 1 centimeter (cm) = 0.01 m

milli m 1/1,000 1 millimeter (mm) = 0.001 m

micro μ 1/1,000,000 1 micrometer (μm) = 10 −6


m

nano n 1/1,000,000,000 1 nanometer (nm) = 10 −9


m

pico p 1/1,000,000,000,000 1 picometer (pm) = 10 −12


m

There are a couple of odd little practices with the use of metric abbreviations. Most abbreviations are lowercase. We use "m" for
meter and not "M ". However, when it comes to volume, the base unit "liter" is abbreviated as "L" and not "l". So we would write
3.5 milliliters as 3.5 mL.
As a practical matter, whenever possible you should express the units in a small and manageable number. If you are measuring the
weight of a material that weighs 6.5 kg, this is easier than saying it weighs 6500 g or 0.65 dag. All three are correct, but the kg
units in this case make for a small and easily managed number. However, if a specific problem needs grams instead of kilograms,
go with the grams for consistency.

1.18.2 https://chem.libretexts.org/@go/page/423581
 Example 1.18.1: Unit Abbreviations

Give the abbreviation for each unit and define the abbreviation in terms of the base unit.
a. kiloliter
b. microsecond
c. decimeter
d. nanogram

Solutions
Solutions to Example 2.5.1
Explanation Answer

The prefix kilo means “1,000 ×,” so 1 kL


a kL
equals 1,000 L.

The prefix micro implies 1/1,000,000th of a


b µs
unit, so 1 µs equals 0.000001 s.

The prefix deci means 1/10th, so 1 dm equals


c dm
0.1 m.

The prefix nano means 1/1000000000, so a


d ng
nanogram is equal to 0.000000001 g.

 Exercise 1.18.1

Give the abbreviation for each unit and define the abbreviation in terms of the base unit.
a. kilometer
b. milligram
c. nanosecond
d. centiliter

Answer a:
km

Answer b:
mg

Answer c:
ns

Answer d:
cL

Summary
Metric prefixes derive from Latin or Greek terms. The prefixes are used to make the units manageable.
The SI system is based on multiples of ten. There are seven basic units in the SI system. Five of these units are commonly used
in chemistry.

1.18: The Basic Units of Measurement is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew &
Henry Agnew.
2.5: The Basic Units of Measurement by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

1.18.3 https://chem.libretexts.org/@go/page/423581
1.19: PROBLEM SOLVING AND UNIT CONVERSIONS
 LEARNING OBJECTIVES
To convert a value reported in one unit to a corresponding value in a different unit using conversion factors.

During your studies of chemistry (and physics also), you will note that mathematical equations are used in many different applications.
Many of these equations have a number of different variables with which you will need to work. You should also note that these equations
will often require you to use measurements with their units. Algebra skills become very important here!

CONVERTING BETWEEN UNITS WITH CONVERSION FACTORS


A conversion factor is a factor used to convert one unit of measurement into another. A simple conversion factor can convert meters into
centimeters, or a more complex one can convert miles per hour into meters per second. Since most calculations require measurements to be
in certain units, you will find many uses for conversion factors. Always remember that a conversion factor has to represent a fact; this fact
can either be simple or more complex. For instance, you already know that 12 eggs equal 1 dozen. A more complex fact is that the speed of
light is 1.86 × 10 miles/sec. Either one of these can be used as a conversion factor depending on what type of calculation you are working
5

with (Table 1.19.1 ).


Table 1.19.1: Conversion Factors from SI units to English Units
English Units Metric Units Quantity
1 ounce (oz) 28.35 grams (g) *mass
1 fluid once (oz) 29.6 mL volume
2.205 pounds (lb) 1 kilogram (kg) *mass
1 inch (in) 2.54 centimeters (cm) length
0.6214 miles (mi) 1 kilometer (km) length
1 quarter (qt) 0.95 liters (L) volume

*Pounds and ounces are technically units of force, not mass, but this fact is often ignored by the non-scientific community.
Of course, there are other ratios which are not listed in Table 1.19.1 . They may include:
Ratios embedded in the text of the problem (using words such as per or in each, or using symbols such as / or %).
Conversions in the metric system, as covered earlier in this chapter.
Common knowledge ratios (such as 60 seconds = 1 minute).
If you learned the SI units and prefixes described, then you know that 1 cm is 1/100th of a meter.
1
−2
1 cm = m = 10 m
100

or
100 cm = 1 m

Suppose we divide both sides of the equation by 1m (both the number and the unit):
100 cm 1 m
=
1 m 1 m

As long as we perform the same operation on both sides of the equals sign, the expression remains an equality. Look at the right side of the
equation; it now has the same quantity in the numerator (the top) as it has in the denominator (the bottom). Any fraction that has the same
quantity in the numerator and the denominator has a value of 1:
6
100 cm 1000 mm 1 × 10 μm
= = = 1
1 m 1 m 1 m

We know that 100 cm is 1 m, so we have the same quantity on the top and the bottom of our fraction, although it is expressed in different
units.

PERFORMING DIMENSIONAL ANALYSIS


Dimensional analysis is amongst the most valuable tools that physical scientists use. Simply put, it is the conversion between an amount in
one unit to the corresponding amount in a desired unit using various conversion factors. This is valuable because certain measurements are

1.19.1 https://chem.libretexts.org/@go/page/423582
more accurate or easier to find than others. The use of units in a calculation to ensure that we obtain the final proper units is called
dimensional analysis.
Here is a simple example. How many centimeters are there in 3.55 m? Perhaps you can determine the answer in your head. If there are 100
cm in every meter, then 3.55 m equals 355 cm. To solve the problem more formally with a conversion factor, we first write the quantity we
are given, 3.55 m. Then we multiply this quantity by a conversion factor, which is the same as multiplying it by 1. We can write 1 as
100 cm
and multiply:
1 m

100 cm
3.55 m ×
1 m

The 3.55 m can be thought of as a fraction with a 1 in the denominator. Because m, the abbreviation for meters, occurs in both the
numerator and the denominator of our expression, they cancel out:
3.55 m 100 cm
×
1 1 m

The final step is to perform the calculation that remains once the units have been canceled:
3.55 100 cm
× = 355 cm
1 1

In the final answer, we omit the 1 in the denominator. Thus, by a more formal procedure, we find that 3.55 m equals 355 cm. A generalized
description of this process is as follows:
quantity (in old units) × conversion factor = quantity (in new units)
You may be wondering why we use a seemingly complicated procedure for a straightforward conversion. In later studies, the conversion
problems you encounter will not always be so simple. If you master the technique of applying conversion factors, you will be able to solve a
large variety of problems.
100 cm 1 m
In the previous example, we used the fraction as a conversion factor. Does the conversion factor also equal 1? Yes, it
1 m 100 cm
does; it has the same quantity in the numerator as in the denominator (except that they are expressed in different units). Why did we not use
that conversion factor? If we had used the second conversion factor, the original unit would not have canceled, and the result would have
been meaningless. Here is what we would have gotten:
2
1 m m
3.55 m × = 0.0355
100 cm cm

For the answer to be meaningful, we have to construct the conversion factor in a form that causes the original unit to cancel out. Figure
1.19.1 shows a concept map for constructing a proper conversion.

Figure 1.19.1: A Concept Map for Conversions. This is how you construct a conversion factor to convert from one unit to another.
Converting meters to centimeters: the unit you are converting is on the bottom of the fraction, and the unit to convert to is on top

 GENERAL STEPS IN PERFORMING DIMENSIONAL ANALYSIS

1. Identify the "given" information in the problem. Look for a number with units to start this problem with.
2. What is the problem asking you to "find"? In other words, what unit will your answer have?
3. Use ratios and conversion factors to cancel out the units that aren't part of your answer, and leave you with units that are part of
your answer.
4. When your units cancel out correctly, you are ready to do the math. You are multiplying fractions, so you multiply the top numbers
and divide by the bottom numbers in the fractions.

1.19.2 https://chem.libretexts.org/@go/page/423582
SIGNIFICANT FIGURES IN CONVERSIONS
How do conversion factors affect the determination of significant figures?
Numbers in conversion factors based on prefix changes, such as kilograms to grams, are not considered in the determination of
significant figures in a calculation because the numbers in such conversion factors are exact.
Exact numbers are defined or counted numbers, not measured numbers, and can be considered as having an infinite number of
significant figures. (In other words, 1 kg is exactly 1,000 g, by the definition of kilo-.)
Counted numbers are also exact. If there are 16 students in a classroom, the number 16 is exact.
In contrast, conversion factors that come from measurements (such as density, as we will see shortly) or that are approximations have a
limited number of significant figures and should be considered in determining the significant figures of the final answer.

 EXAMPLE 1.19.1

Steps for Problem Solving for Example 2.6.1 and 2.6.2


Example 1.19.1 Example 1.19.2
The average volume of blood in an adult male is 4.7 L. A hummingbird can flap its wings once in 18 ms. How
Steps for Problem Solving
What is this volume in milliliters? many seconds are in 18 ms?
Identify the "given" information and what the Given: 4.7 L Given: 18 ms
problem is asking you to "find." Find: mL Find: s
List other known quantities. 1 mL = 10
−3
L 1 ms = 10
−3
s

Prepare a concept map and use the proper


conversion factor.

1 mL
4.7 L × = 4, 700 mL
−3 −3
10 L 10 s
18 ms × = 0.018 s
or 1 ms

Cancel units and calculate. 1, 000 mL


or
4.7 L × = 4, 700 mL
1 s
1 L
18 ms × = 0.018 s
or 1, 000 ms

4.7 x 103 2SF, not ambiguous


The amount in mL should be 1000 times larger than the The amount in s should be 1/1000 the given amount in
Think about your result.
given amount in L. ms.

 EXERCISE 1.19.1

Perform each conversion.


a. 101,000 ns to seconds
b. 32.08 kg to grams
c. 1.53 grams to cg

Answer a:
−4
1.01000x10 s

Answer b:
4
3.208x10 g

Answer c:
2
1.53x10 cg

SUMMARY
Conversion factors are used to convert one unit of measurement into another.
Dimensional analysis (unit conversions) involves the use of conversion factors that will cancel unwanted units and produce the
appropriate units.

1.19: Problem Solving and Unit Conversions is shared under a CC BY-NC-SA 3.0 license and was authored, remixed, and/or curated by Marisa Alviar-
Agnew & Henry Agnew.
2.6: Problem Solving and Unit Conversions by Henry Agnew, Marisa Alviar-Agnew is licensed CC BY-NC-SA 3.0.

1.19.3 https://chem.libretexts.org/@go/page/423582
1.20: SOLVING MULTI-STEP CONVERSION PROBLEMS
MULTIPLE CONVERSIONS
Sometimes you will have to perform more than one conversion to obtain the desired unit. For example, suppose you want to convert 54.7
km into millimeters. We will set up a series of conversion factors so that each conversion factor produces the next unit in the sequence. We
first convert the given amount in km to the base unit, which is meters. We know that 1,000 m =1 km.
Then we convert meters to mm, remembering that 1 mm = 10 −3
m .
CONCEPT MAP

Convert kilometers to meters to millimeters: use conversion factors 1000 meters per 1 kilometer and 1 millimeter per 0.001 meter
CALCULATION

1, 000 m 1 mm
54.7 km × × = 54, 700, 000 mm
1 km −3
10 m

7
= 5.47 × 10 mm

In each step, the previous unit is canceled and the next unit in the sequence is produced, each
successive unit canceling out until only the unit needed in the answer is left.

 EXAMPLE 1.20.1: UNIT CONVERSION


Convert 58.2 ms to megaseconds in one multi-step calculation.
SOLUTION
Solution for Example 2.7.1
Steps for Problem Solving Unit Conversion

Identify the "given" information and what the problem is Given: 58.2 ms
asking you to "find." Find: Ms
−3
1ms = 10 s
List other known quantities 6
1M s = 10 s

Prepare a concept map.

Convert milliseconds to seconds to microseconds: use conversion factors 0.001 second per millisecond
and 1 microsecond per 1 million seconds

−3
10 s 1 Ms
58.2 ms × × = 0.0000000582 Ms
1 ms 1, 000, 000 s
Calculate. −8
= 5.82 × 10 Ms

Neither conversion factor affects the number of significant figures in the final answer.

1.20.1 https://chem.libretexts.org/@go/page/423583
 EXAMPLE 1.20.2: UNIT CONVERSION

How many seconds are in a day?

Solution
Solution for Example 2.7.2
Steps for Problem Solving Unit Conversion

Identify the "given" information and what the Given: 1 day


problem is asking you to "find." Find: s
1 day = 24 hours
List other known quantities. 1 hour = 60 minutes
1 minute = 60 seconds

Prepare a concept map.

Convert day to hour to minute to second: use conversion factors 24 hours per day, 60 minutes per hour, and 60
seconds per minute

24 hr 60 min 60 s
Calculate. 1 d × × × = 86, 400 s
1 d 1 hr 1 min

 EXERCISE 1.20.1

Perform each conversion in one multi-step calculation.


a. 43.007 ng to kg
b. 1005 in to ft
c. 12 mi to km

Answer a
−11
4.3007 × 10 kg

Answer b
83.75 ft

Answer c
19 km

 CAREER FOCUS: PHARMACIST


A pharmacist dispenses drugs that have been prescribed by a doctor. Although that may sound straightforward, pharmacists in the
United States must hold a doctorate in pharmacy and be licensed by the state in which they work. Most pharmacy programs require
four years of education in a specialty pharmacy school. Pharmacists must know a lot of chemistry and biology so they can understand
the effects that drugs (which are chemicals, after all) have on the body. Pharmacists can advise physicians on the selection, dosage,
interactions, and side effects of drugs. They can also advise patients on the proper use of their medications, including when and how to
take specific drugs properly. Pharmacists can be found in drugstores, hospitals, and other medical facilities. Curiously, an outdated
name for pharmacist is chemist, which was used when pharmacists formerly did a lot of drug preparation, or compounding. In modern
times, pharmacists rarely compound their own drugs, but their knowledge of the sciences, including chemistry, helps them provide
valuable services in support of everyone’s health.

1.20.2 https://chem.libretexts.org/@go/page/423583
A woman consulting with a pharmacist. (Public Domain; Rhoda Baer via National Cancer Institute, an agency that is part of the
National Institutes of Health.)

SUMMARY
In multi-step conversion problems, the previous unit is canceled for each step and the next unit in the sequence is produced, each successive
unit canceling out until only the unit needed in the answer is left.

1.20: Solving Multi-step Conversion Problems is shared under a CC BY-NC-SA 3.0 license and was authored, remixed, and/or curated by Marisa Alviar-
Agnew & Henry Agnew.
2.7: Solving Multi-step Conversion Problems by Henry Agnew, Marisa Alviar-Agnew is licensed CC BY-NC-SA 3.0.

1.20.3 https://chem.libretexts.org/@go/page/423583
1.21: Units Raised to a Power
 Learning Objectives
To convert a value reported in one unit raised to a power of 10, to a corresponding value in a different unit raised to the
same power of 10, using conversion factors.

Conversion factors for area and volume can also be produced by the dimensional analysis method. Just remember that if a quantity
is raised to a power of 10, both the number and the unit must be raised to the same power of 10. For example, to convert 1500 cm 2

to m , we need to start with the relationship between centimeter and meter. We know that 1 cm = 10-2 m or 100 cm =1 m, but since
2

we are given the quantity in 1500 cm2, then we have to use the relationship:
2 −2 2 −4 2
1 cm = (10 m) = 10 m

CONCEPT MAP

To convert centimeters squared to meters squared, use the conversion factor 0.01 meters per 1 centimeter, squared overall

CALCULATION
−2 2

2
10 m 2
1500 cm ×( ) = 0.15 m
1 cm

or
2

2
1 m
2
1500 cm ×( ) = 0.15 m
100 cm

or
2
2
1 m 2
1500 cm × = 0.15 m
2
10, 000 cm

 Example 1.21.1: Volume of a Sphere


What is the volume of a sphere (radius 4.30 inches) in cubic cm (cm3)?

Solution
Solution for Example 2.8.1
What is the volume of a sphere (radius 4.30 inches) in cubic cm
Steps for Problem Solving
(cm3)?

Identify the "given” information and what the problem is asking you Given: radius = 4.30 in
to "find." Find: cm3 (volume)
4
Volume of a sphere: V = 3
× π× r
3

Determine other known quantities. =


4
× 3.1416 × (4.3 0in )
3

3 –

= 333

.04in
3

1.21.1 https://chem.libretexts.org/@go/page/423584
What is the volume of a sphere (radius 4.30 inches) in cubic cm
Steps for Problem Solving
(cm3)?

Prepare a concept map.

To convert inches cubed to centimeters cubed, use conversion factor


2.54 centimeters per 1 inch, cubed overall
3

2.54cm
Calculate. 33 3.04 in
3
( )
3
= 5.46 × 10 cm
3


1 in

A centimeter is a smaller unit than an inch, so the answer in cubic


Think about your result.
centimeters is larger than the given value in cubic inches.

 Exercise 1.21.1

Lake Tahoe has a surface area of 191 square miles. What is the area in square km (km2)?

Answer
495 km2

Contributions & Attributions

1.21: Units Raised to a Power is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew & Henry
Agnew.
2.8: Units Raised to a Power by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

1.21.2 https://chem.libretexts.org/@go/page/423584
1.22: Density
 Learning Objectives
Define density.
Use density as a conversion factor.

Density (ρ) is a physical property found by dividing the mass of an object by its volume. Regardless of the sample size, density is
always constant. For example, the density of a pure sample of tungsten is always 19.25 grams per cubic centimeter. This means that
whether you have one gram or one kilogram of the sample, the density will never vary. The equation, as we already know, is as
follows:
Mass
Density =
Volume

or just
m
ρ = (1.22.1)
V

Based on this equation, it's clear that density can, and does, vary from element to element and substance to substance due to
differences in the relationship of mass and volume. Pure water, for example, has a density of 0.998 g/cm3 at 25° C. The average
densities of some common substances are in Table 1.22.1. Notice that corn oil has a lower mass to volume ratio than water. This
means that when added to water, corn oil will “float.”
Table 1.22.1 : Densities of Common Substances
Substance Density at 25°C (g/cm3)

blood 1.035

body fat 0.918

whole milk 1.030

corn oil 0.922

mayonnaise 0.910

honey 1.420

Density can be measured for all substances—solids, liquids and gases. For solids and liquids, density is often reported using the
units of g/cm3. Densities of gases, which are significantly lower than the densities of solids and liquids, are often given using units
of g/L.

 Example 1.22.1: Ethyl Alcohol


Calculate the density of a 30.2 mL sample of ethyl alcohol with a mass of 23.71002 g

Solution
This is a direct application of Equation 1.22.1:
23.71002 g
ρ = = 0.785 g/mL
30.2 mL

 Exercise 1.22.1
a. Find the density (in kg/L) of a sample that has a volume of 36.5 L and a mass of 10.0 kg.
b. If you have a 2.130 mL sample of acetic acid with mass 0.002234 kg, what is the density in kg/L?

Answer a
0.274 kg/L

1.22.1 https://chem.libretexts.org/@go/page/423585
Answer b
1.049 kg/L

Density as a Conversion Factor


Conversion factors can also be constructed for converting between different kinds of units. For example, density can be used to
convert between the mass and the volume of a substance. Consider mercury, which is a liquid at room temperature and has a
density of 13.6 g/mL. The density tells us that 13.6 g of mercury have a volume of 1 mL. We can write that relationship as follows:
13.6 g mercury = 1 mL mercury
This relationship can be used to construct two conversion factors:
13.6 g
=1
1 mL

and
1 mL
=1
13.6 g

Which one do we use? It depends, as usual, on the units we need to cancel and introduce. For example, suppose we want to know
the mass of 2.0 mL of mercury. We would use the conversion factor that has milliliters on the bottom (so that the milliliter unit
cancels) and grams on top, so that our final answer has a unit of mass:
13.6 g
2.0 mL × = 27.2 g = 27 g
1 mL

In the last step, we limit our final answer to two significant figures because the volume quantity has only two significant figures;
the 1 in the volume unit is considered an exact number, so it does not affect the number of significant figures. The other conversion
factor would be useful if we were given a mass and asked to find volume, as the following example illustrates.

Density can be used as a conversion factor between mass and volume.

 Example 1.22.2: Mercury Thermometer Steps for Problem Solving

A mercury thermometer for measuring a patient’s temperature contains 0.750 g of mercury. What is the volume of this mass of
mercury?

Solution
Solution to Example 2.9.2
Steps for Problem Solving Unit Conversion

Identify the "given" information and what the problem is asking you Given: 0.750 g
to "find." Find: mL

List other known quantities. 13.6 g/mL (density of mercury)

Prepare a concept map.

1 mL
0.750 g × = 0.055147. . . mL ≈ 0.0551 mL
Calculate. 13.6 g

We have limited the final answer to three significant figures.

1.22.2 https://chem.libretexts.org/@go/page/423585
 Exercise 1.22.2

What is the volume of 100.0 g of air if its density is 1.3 g/L?

Answer
77 L

Summary
Density is defined as the mass of an object divided by its volume.
Density can be used as a conversion factor between mass and volume.

Contributions & Attributions


1.22: Density is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew & Henry Agnew.
2.9: Density by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source: https://www.ck12.org/c/chemistry/.

1.22.3 https://chem.libretexts.org/@go/page/423585
CHAPTER OVERVIEW

2: Module 2 - Atoms and Bonding


2.1: Indivisible - The Atomic Theory
2.2: The Nuclear Atom
2.3: The Properties of Protons, Neutrons, and Electrons
2.4: Elements- Defined by Their Number of Protons
2.5: Elements- Defined by Their Number of Protons
2.6: Looking for Patterns - The Periodic Table
2.7: Ions - Losing and Gaining Electrons
2.8: Isotopes - When the Number of Neutrons Varies
2.9: Atomic Mass - The Average Mass of an Element’s Atoms
2.10: Light- Electromagnetic Radiation
2.11: The Electromagnetic Spectrum
2.12: The Bohr Model - Atoms with Orbits
2.13: The Quantum-Mechanical Model- Atoms with Orbitals
2.14: Quantum-Mechanical Orbitals and Electron Configurations
2.15: The Explanatory Power of the Quantum-Mechanical Model
2.16: Periodic Trends - Atomic Size, Ionization Energy, and Metallic Character
2.17: Electron Configurations and the Periodic Table
2.18: Representing Valence Electrons with Dots
2.19: Lewis Structures of Ionic Compounds- Electrons Transferred
2.20: Covalent Lewis Structures- Electrons Shared
2.21: Writing Lewis Structures for Covalent Compounds
2.22: Predicting the Shapes of Molecules
2.23: Electronegativity and Polarity - Why Oil and Water Do not Mix
2.24: A Molecular View of Elements and Compounds
2.25: Writing Formulas for Ionic Compounds
2.26: Nomenclature- Naming Compounds
2.27: Naming Ionic Compounds
2.28: Naming Molecular Compounds

2: Module 2 - Atoms and Bonding is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.

1
2.1: Indivisible - The Atomic Theory
 Learning Objectives
Give a short history of the concept of the atom.
Describe the contributions of Democritus and Dalton to atomic theory.
Summarize Dalton's atomic theory and explain its historical development.

You learned earlier that all matter in the universe is made out of tiny building blocks called atoms. All modern scientists accept the
concept of the atom, but when the concept of the atom was first proposed about 2,500 years ago, ancient philosophers laughed at
the idea. It has always been difficult to convince people of the existence of things that are too small to see. We will spend some
time considering the evidence (observations) that convince scientists of the existence of atoms.

Democritus and the Greek Philosophers


About 2,500 years ago, early Greek philosophers believed the entire universe was a single, huge, entity. In other words, "everything
was one." They believed that all objects, all matter, and all substances were connected as a single, big, unchangeable "thing." One
of the first people to propose "atoms" was a man known as Democritus. As an alternative to the beliefs of the Greek philosophers,
he suggested that atomos, or atomon—tiny, indivisible, solid objects—make up all matter in the universe.

Figure 2.1.1 : (left) Democritus by Hendrick ter Brugghen, 1628. Democritus was known as the "laughing philosopher." It was a
good thing he liked to laugh, because most other philosophers were laughing at his theories. (right) British physicist and chemist
John Dalton (1766-1844). Unlike the Greek philosophers, John Dalton believed in both logical thinking and experimentation.
Democritus then reasoned that changes occur when the many atomos in an object were reconnected or recombined in different
ways. Democritus even extended this theory, suggesting that there were different varieties of atomos with different shapes, sizes,
and masses. He thought, however, that shape, size, and mass were the only properties differentiating the different types of atomos.
According to Democritus, other characteristics, like color and taste, did not reflect properties of the atomos themselves, but rather,
resulted from the different ways in which the atomos were combined and connected to one another.
The early Greek philosophers tried to understand the nature of the world through reason and logic, but not through experiment and
observation. As a result, they had some very interesting ideas, but they felt no need to justify their ideas based on life experiences.
In a lot of ways, you can think of the Greek philosophers as being "all thought and no action." It's truly amazing how much they
achieved using their minds, but because they never performed any experiments, they missed or rejected a lot of discoveries that
they could have made otherwise. Greek philosophers dismissed Democritus' theory entirely. Sadly, it took over two millennia
before the theory of atomos (or "atoms," as they are known today) was fully appreciated.

 Greeks: "All Thought and No Action"

Greek philosophers were "all thought and no action" and did not feel the need to test their theories with reality. In contrast,
Dalton's efforts were based on experimentation and testing ideas against reality.

While it must be assumed that many more scientists, philosophers, and others studied composition of matter after Democritus, a
major leap forward in our understanding of the composition of matter took place in the 1800's with the work of the British
scientists John Dalton. He started teaching school at age twelve, and was primarily known as a teacher. In his twenties, he moved to

2.1.1 https://chem.libretexts.org/@go/page/423587
the growing city of Manchester, where he was able to pursue some scientific studies. His work in several areas of science brought
him a number of honors. When he died, over 40,000 people in Manchester marched at his funeral.
The modern atomic theory, proposed about 1803 by the English chemist John Dalton (Figure 2.1.1), is a fundamental concept that
states that all elements are composed of atoms. Previously, we defined an atom as the smallest part of an element that maintains the
identity of that element. Individual atoms are extremely small; even the largest atom has an approximate diameter of only 5.4 ×
10−10 m. With that size, it takes over 18 million of these atoms, lined up side by side, to equal the width of your little finger (about
1 cm).
Dalton studied the weights of various elements and compounds. He noticed that matter always combined in fixed ratios based on
weight, or volume in the case of gases. Chemical compounds always contain the same proportion of elements by mass, regardless
of amount, which provided further support for Proust's law of definite proportions. Dalton also observed that there could be more
than one combination of two elements.
From his experiments and observations, as well as the work from peers of his time, Dalton proposed a new theory of the atom. This
later became known as Dalton's atomic theory. The general tenets of this theory were as follows:
All matter is composed of extremely small particles called atoms.
Atoms of a given element are identical in size, mass, and other properties. Atoms of different elements differ in size, mass, and
other properties.
Atoms cannot be subdivided, created, or destroyed.
Atoms of different elements can combine in simple whole number ratios to form chemical compounds.
In chemical reactions, atoms are combined, separated, or rearranged.
Dalton's atomic theory has been largely accepted by the scientific community, with the exception of three changes. We know now
that (1) an atom can be further subdivided, (2) all atoms of an element are not identical in mass, and (3) using nuclear fission and
fusion techniques, we can create or destroy atoms by changing them into other atoms.

Figure 2.1.2 : Dalton's symbols from his text "A New System of Chemical Philosophy."
The evidence for atoms is so great that few doubt their existence. In fact, individual atoms are now routinely observed with state-
of-the art technologies. Moreover, they can even be used for making pretty images; or as IBM research demonstrates in Video
2.1.1, control of individual atoms can be use used create animations.

2.1.2 https://chem.libretexts.org/@go/page/423587
A Boy And His Atom: The World's Small…
Small…

Video 2.1.1 : A Boy And His Atom: The World's Smallest Movie.
A Boy and His Atom is a 2012 stop-motion animated short film released by IBM Research. The movie tells the story of a boy and a
wayward atom who meet and become friends. It depicts a boy playing with an atom that takes various forms. It was made by
moving carbon monoxide molecules viewed with a scanning tunneling microscope, a device that magnifies them 100 million times.
These molecules were moved to create images, which were then saved as individual frames to make the film.

Summary
2,500 years ago, Democritus suggested that all matter in the universe was made up of tiny, indivisible, solid objects he called
"atomos." However, other Greek philosophers disliked Democritus' "atomos" theory because they felt it was illogical.
Dalton's Atomic Theory is the first scientific theory to relate chemical changes to the structure, properties, and behavior of the
atom. The general tenets of this theory are:
All matter is composed of extremely small particles called atoms.
Atoms of a given element are identical in size, mass, and other properties. Atoms of different elements differ in size, mass,
and other properties.
Atoms cannot be subdivided, created, or destroyed.
Atoms of different elements can combine in simple whole number ratios to form chemical compounds.
In chemical reactions, atoms are combined, separated, or rearranged.

2.1: Indivisible - The Atomic Theory is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew &
Henry Agnew.
4.2: Indivisible - The Atomic Theory by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

2.1.3 https://chem.libretexts.org/@go/page/423587
2.2: The Nuclear Atom
 Learning Objectives
Explain the observations that led to Thomson's discovery of the electron.
Describe Thomson's "plum pudding" model of the atom and the evidence for it.
Draw a diagram of Thomson's "plum pudding" model of the atom and explain why it has this name.
Describe Rutherford's gold foil experiment and explain how this experiment altered the "plum pudding" model.
Draw a diagram of the Rutherford model of the atom and label the nucleus and the electron cloud.

Dalton's Atomic Theory held up well to a lot of the different chemical experiments that scientists performed to test it. In fact, for
almost 100 years, it seemed as if Dalton's Atomic Theory was the whole truth. However, in 1897, a scientist named J. J. Thomson
conducted some research that suggested that Dalton's Atomic Theory was not the entire story. He suggested that the small,
negatively charged particles making up the cathode ray were actually pieces of atoms. He called these pieces "corpuscles,"
although today we know them as electrons. Thanks to his clever experiments and careful reasoning, J. J. Thomson is credited with
the discovery of the electron.

Figure 2.2.1 : J. J. Thomson (center) concluded experiments that suggested that Dalton's (left) atomic theory was not telling the
entire story. Ernest Rutherford (right) later built on Thomson's work to argue for the model nuclear atom.

Electrons and Plums


The electron was discovered by J. J. Thomson in 1897. The existence of protons was also known, as was the fact that atoms were
neutral in charge. Since the intact atom had no net charge and the electron and proton had opposite charges, the next step after the
discovery of subatomic particles was to figure out how these particles were arranged in the atom. This was a difficult task because
of the incredibly small size of the atom. Therefore, scientists set out to design a model of what they believed the atom could look
like. The goal of each atomic model was to accurately represent all of the experimental evidence about atoms in the simplest way
possible.
Following the discovery of the electron, J.J. Thomson developed what became known as the "plum pudding" model in 1904. Plum
pudding is an English dessert similar to a blueberry muffin. In Thomson's plum pudding model of the atom, the electrons were
embedded in a uniform sphere of positive charge like blueberries stuck into a muffin. The positive matter was thought to be jelly-
like or similar to a thick soup. The electrons were somewhat mobile. As they got closer to the outer portion of the atom, the
positive charge in the region was greater than the neighboring negative charges, and the electron would be pulled back more toward
the center region of the atom.

Figure 2.2.2 : The "plum pudding" model.

2.2.1 https://chem.libretexts.org/@go/page/423588
However, this model of the atom soon gave way to a new model developed by New Zealander Ernest Rutherford (1871-1937)
about five years later. Thomson did still receive many honors during his lifetime, including being awarded the Nobel Prize in
Physics in 1906 and a knighthood in 1908.

Atoms and Gold


In 1911, Rutherford and coworkers Hans Geiger and Ernest Marsden initiated a series of groundbreaking experiments that would
completely change the accepted model of the atom. They bombarded very thin sheets of gold foil with fast moving alpha particles.
Alpha particles, a type of natural radioactive particle, are positively charged particles with a mass about four times that of a
hydrogen atom.

Figure 2.2.3 : (A) The experimental setup for Rutherford's gold foil experiment: A radioactive element that emitted alpha particles
was directed toward a thin sheet of gold foil that was surrounded by a screen which would allow detection of the deflected
particles. (B) According to the plum pudding model (top), all of the alpha particles should have passed through the gold foil with
little or no deflection. Rutherford found that a small percentage of alpha particles were deflected at large angles, which could be
explained by an atom with a very small, dense, positively-charged nucleus at its center (bottom).
According to the accepted atomic model, in which an atom's mass and charge are uniformly distributed throughout the atom, the
scientists expected that all of the alpha particles would pass through the gold foil with only a slight deflection or none at all.
Surprisingly, while most of the alpha particles were indeed not deflected, a very small percentage (about 1 in 8000 particles)
bounced off the gold foil at very large angles. Some were even redirected back toward the source. No prior knowledge had
prepared them for this discovery. In a famous quote, Rutherford exclaimed that it was "as if you had fired a 15-inch [artillery] shell
at a piece of tissue and it came back and hit you."
Rutherford needed to come up with an entirely new model of the atom in order to explain his results. Because the vast majority of
the alpha particles had passed through the gold, he reasoned that most of the atom was empty space. In contrast, the particles that
were highly deflected must have experienced a tremendously powerful force within the atom. He concluded that all of the positive
charge and the majority of the mass of the atom must be concentrated in a very small space in the atom's interior, which he called
the nucleus. The nucleus is the tiny, dense, central core of the atom and is composed of protons and neutrons.
Rutherford's atomic model became known as the nuclear model. In the nuclear atom, the protons and neutrons, which comprise
nearly all of the mass of the atom, are located in the nucleus at the center of the atom. The electrons are distributed around the
nucleus and occupy most of the volume of the atom. It is worth emphasizing just how small the nucleus is compared to the rest of
the atom. If we could blow up an atom to be the size of a large professional football stadium, the nucleus would be about the size of
a marble.
Rutherford's model proved to be an important step towards a full understanding of the atom. However, it did not completely
address the nature of the electrons and the way in which they occupy the vast space around the nucleus. It was not until some years
later that a full understanding of the electron was achieved. This proved to be the key to understanding the chemical properties of
elements.

Atomic Nucleus
The nucleus (plural, nuclei) is a positively charged region at the center of the atom. It consists of two types of subatomic particles
packed tightly together. The particles are protons, which have a positive electric charge, and neutrons, which are neutral in electric
charge. Outside of the nucleus, an atom is mostly empty space, with orbiting negative particles called electrons whizzing through it.
The figure below shows these parts of the atom.

2.2.2 https://chem.libretexts.org/@go/page/423588
Figure 2.2.4 : The nuclear atom.
The nucleus of the atom is extremely small. Its radius is only about 1/100,000 of the total radius of the atom. Electrons have
virtually no mass, but protons and neutrons have a lot of mass for their size. As a result, the nucleus has virtually all the mass of an
atom. Given its great mass and tiny size, the nucleus is very dense. If an object the size of a penny had the same density as the
nucleus of an atom, its mass would be greater than 30 million tons!

 Holding it all Together

Particles with opposite electric charges attract each other. This explains why negative electrons orbit the positive nucleus.
Particles with the same electric charge repel each other. This means that the positive protons in the nucleus push apart from one
another. So why doesn't the nucleus fly apart? An even stronger force—called the strong nuclear force—holds protons and
neutrons together in the nucleus.

Summary
Atoms are the ultimate building blocks of all matter.
The modern atomic theory establishes the concepts of atoms and how they compose matter.
Bombardment of gold foil with alpha particles showed that some particles were deflected.
The nuclear model of the atom consists of a small and dense positively charged interior surrounded by a cloud of electrons.

This page titled 2.2: The Nuclear Atom is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew &
Henry Agnew via source content that was edited to the style and standards of the LibreTexts platform; a detailed edit history is available upon
request.
4.3: The Nuclear Atom by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source: https://www.ck12.org/c/chemistry/.

2.2.3 https://chem.libretexts.org/@go/page/423588
2.3: The Properties of Protons, Neutrons, and Electrons
 Learning Objectives
Describe the locations, charges, and masses of the three main subatomic particles.
Determine the number of protons and electrons in an atom.
Define atomic mass unit (amu).

Dalton's Atomic Theory explained a lot about matter, chemicals, and chemical reactions. Nevertheless, it was not entirely accurate,
because contrary to what Dalton believed, atoms can, in fact, be broken apart into smaller subunits or subatomic particles. We have
been talking about the electron in great detail, but there are two other particles of interest to us: protons and neutrons. We already
learned that J. J. Thomson discovered a negatively charged particle, called the electron. Rutherford proposed that these electrons
orbit a positive nucleus. In subsequent experiments, he found that there is a smaller positively charged particle in the nucleus,
called a proton. There is also a third subatomic particle, known as a neutron.

Electrons
Electrons are one of three main types of particles that make up atoms. Unlike protons and neutrons, which consist of smaller,
simpler particles, electrons are fundamental particles that do not consist of smaller particles. They are a type of fundamental
particle called leptons. All leptons have an electric charge of −1 or 0. Electrons are extremely small. The mass of an electron is
only about 1/2000 the mass of a proton or neutron, so electrons contribute virtually nothing to the total mass of an atom. Electrons
have an electric charge of −1, which is equal but opposite to the charge of a proton, which is +1. All atoms have the same number
of electrons as protons, so the positive and negative charges "cancel out", making atoms electrically neutral.
Unlike protons and neutrons, which are located inside the nucleus at the center of the atom, electrons are found outside the nucleus.
Because opposite electric charges attract one another, negative electrons are attracted to the positive nucleus. This force of
attraction keeps electrons constantly moving through the otherwise empty space around the nucleus. The figure below is a common
way to represent the structure of an atom. It shows the electron as a particle orbiting the nucleus, similar to the way that planets
orbit the sun. However, this is an incorrect perspective, as quantum mechanics demonstrates that electrons are more complicated.

Figure 2.3.1 : Electrons are much smaller than protons or neutrons. If an electron was the mass of a penny, a proton or a neutron
would have the mass of a large bowling ball!

Protons
A proton is one of three main particles that make up the atom. Protons are found in the nucleus of the atom. This is a tiny, dense
region at the center of the atom. Protons have a positive electrical charge of one (+1) and a mass of 1 atomic mass unit (amu),
which is about 1.67 × 10 −27
kilograms. Together with neutrons, they make up virtually all of the mass of an atom.

Neutrons
Atoms of all elements—except for most atoms of hydrogen—have neutrons in their nucleus. Unlike protons and electrons, which
are electrically charged, neutrons have no charge—they are electrically neutral. That's why the neutrons in the diagram above are
labeled n . The zero stands for "zero charge". The mass of a neutron is slightly greater than the mass of a proton, which is 1 atomic
0

mass unit (amu). (An atomic mass unit equals about 1.67 × 10 kilograms.) A neutron also has about the same diameter as a
−27

proton, or 1.7 × 10−15


meters.

2.3.1 https://chem.libretexts.org/@go/page/423589
As you might have already guessed from its name, the neutron is neutral. In other words, it has no charge whatsoever and is
therefore neither attracted to nor repelled from other objects. Neutrons are in every atom (with one exception), and they are bound
together with other neutrons and protons in the atomic nucleus.
Before we move on, we must discuss how the different types of subatomic particles interact with each other. When it comes to
neutrons, the answer is obvious. Since neutrons are neither attracted to nor repelled from objects, they don't really interact with
protons or electrons (beyond being bound into the nucleus with the protons).
Even though electrons, protons, and neutrons are all types of subatomic particles, they are not all the same size. When you compare
the masses of electrons, protons, and neutrons, what you find is that electrons have an extremely small mass, compared to either
protons or neutrons. On the other hand, the masses of protons and neutrons are fairly similar, although technically, the mass of a
neutron is slightly larger than the mass of a proton. Because protons and neutrons are so much more massive than electrons, almost
all of the mass of any atom comes from the nucleus, which contains all of the neutrons and protons.
Table 2.3.1 : Properties of Subatomic Particles
Relative Mass
Particle Symbol Mass (amu) Relative Charge Location
(proton = 1)

proton p+ 1 1 +1 inside the nucleus

electron e− 5.45 × 10−4 0.00055 −1 outside the nucleus

neutron n0 1 1 0 inside the nucleus

Table 2.3.1 gives the properties and locations of electrons, protons, and neutrons. The third column shows the masses of the three
subatomic particles in "atomic mass units." An atomic mass unit (amu) is defined as one-twelfth of the mass of a carbon-12 atom.
Atomic mass units (amu) are useful, because, as you can see, the mass of a proton and the mass of a neutron are almost exactly 1
in this unit system.
Negative and positive charges of equal magnitude cancel each other out. This means that the negative charge on an electron
perfectly balances the positive charge on the proton. In other words, a neutral atom must have exactly one electron for every
proton. If a neutral atom has 1 proton, it must have 1 electron. If a neutral atom has 2 protons, it must have 2 electrons. If a neutral
atom has 10 protons, it must have 10 electrons. You get the idea. In order to be neutral, an atom must have the same number of
electrons and protons.

Summary
Electrons are a type of subatomic particle with a negative charge.
Protons are a type of subatomic particle with a positive charge. Protons are bound together in an atom's nucleus as a result of
the strong nuclear force.
Neutrons are a type of subatomic particle with no charge (they are neutral). Like protons, neutrons are bound into the atom's
nucleus as a result of the strong nuclear force.
Protons and neutrons have approximately the same mass, but they are both much more massive than electrons (approximately
2,000 times as massive as an electron).
The positive charge on a proton is equal in magnitude to the negative charge on an electron. As a result, a neutral atom must
have an equal number of protons and electrons.
The atomic mass unit (amu) is a unit of mass equal to one-twelfth the mass of a carbon-12 atom

2.3: The Properties of Protons, Neutrons, and Electrons is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa
Alviar-Agnew & Henry Agnew.
4.4: The Properties of Protons, Neutrons, and Electrons by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

2.3.2 https://chem.libretexts.org/@go/page/423589
2.4: Elements- Defined by Their Number of Protons
 Learning Objectives
Define atomic number.
Define mass number.
Determine the number of protons, neutrons, and electrons in an atom.

It's important to be able to distinguish atoms of one element from atoms of another element. Elements are pure substances that
make up all other matter, so each one is given a unique name. The names of elements are also represented by unique one- or two-
letter symbols, such as H for hydrogen, C for carbon, or He for helium. However, it would more powerful if these names could be
used to identify the numbers of protons and neutrons in the atoms. That's where atomic number and mass number are useful.

Figure 2.4.1 : It is difficult to find qualities that differ between each element, and to distinguish one element from another. Each
element, however, does have a unique number of protons. Sulfur has 16 protons, silicon has 14 protons, and gold has 79 protons.
Images used with permission (public domain for sulfur and silicon, gold is licensed by CC-BY-SA-NC-ND; Alchemist-hp).

Atomic Number
Scientists distinguish between different elements by counting the number of protons in the nucleus (Table 2.4.1). If an atom has
only one proton, we know that it's a hydrogen atom. An atom with two protons is always a helium atom. If scientists count four
protons in an atom, they know it's a beryllium atom. An atom with three protons is a lithium atom, an atom with five protons is a
boron atom, an atom with six protons is a carbon atom . . . the list goes on.
Since an atom of one element can be distinguished from an atom of another element by the number of protons in its nucleus,
scientists are always interested in this number, and how this number differs between different elements. The number of protons in
an atom is called its atomic number (Z ). This number is very important because it is unique for atoms of a given element. All
atoms of an element have the same number of protons, and every element has a different number of protons in its atoms. For
example, all helium atoms have two protons, and no other elements have atoms with two protons.
Table 2.4.1 : Atoms of the First Six Elements
Name Protons Neutrons Electrons Atomic Number (Z) Mass Number (A)

Hydrogen 1 0 1 1 1

Helium 2 2 2 2 4

Lithium 3 4 3 3 7

Beryllium 4 5 4 4 9

Boron 5 6 5 5 11

Carbon 6 6 6 6 12

Of course, since neutral atoms have to have one electron for every proton, an element's atomic number also tells you how many
electrons are in a neutral atom of that element. For example, hydrogen has an atomic number of 1. This means that an atom of
hydrogen has one proton, and, if it's neutral, one electron as well. Gold, on the other hand, has an atomic number of 79, which
means that an atom of gold has 79 protons, and, if it's neutral, 79 electrons as well.

2.4.1 https://chem.libretexts.org/@go/page/423590
 Neutral Atoms

Atoms are neutral in electrical charge because they have the same number of negative electrons as positive protons (Table
2.4.1). Therefore, the atomic number of an atom also tells you how many electrons the atom has. This, in turn, determines

many of the atom's chemical properties.

Mass Number
The mass number (A ) of an atom is the total number of protons and neutrons in its nucleus. The mass of the atom is a unit called
the atomic mass unit (amu). One atomic mass unit is the mass of a proton, or about 1.67 × 10 −27
kilograms, which is an
extremely small mass. A neutron has just a tiny bit more mass than a proton, but its mass is often assumed to be one atomic mass
unit as well. Because electrons have virtually no mass, just about all the mass of an atom is in its protons and neutrons. Therefore,
the total number of protons and neutrons in an atom determines its mass in atomic mass units (Table 2.4.1).
Consider helium again. Most helium atoms have two neutrons in addition to two protons. Therefore the mass of most helium atoms
is 4 atomic mass units (2 amu for the protons + 2 amu for the neutrons). However, some helium atoms have more or less than two
neutrons. Atoms with the same number of protons but different numbers of neutrons are called isotopes. Because the number of
neutrons can vary for a given element, the mass numbers of different atoms of an element may also vary. For example, some
helium atoms have three neutrons instead of two (these are called isotopes and are discussed in detail later on).
Why do you think that the "mass number" includes protons and neutrons, but not electrons? You know that most of the mass of an
atom is concentrated in its nucleus. The mass of an atom depends on the number of protons and neutrons. You have already learned
that the mass of an electron is very, very small compared to the mass of either a proton or a neutron (like the mass of a penny
compared to the mass of a bowling ball). Counting the number of protons and neutrons tells scientists about the total mass of an
atom.

mass number A = (number of protons) + (number of neutrons)

An atom's mass number is very easy to calculate, provided that you know the number of protons and neutrons in an atom.

 Example 4.5.1

What is the mass number of an atom of helium that contains 2 neutrons?

Solution
(number of protons) = 2 (Remember that an atom of helium always has 2 protons.)
(number of neutrons) = 2

mass number = (number of protons) + (number of neutrons)

mass number = 2 + 2 = 4

A chemical symbol is a one- or two-letter designation of an element. Some examples of chemical symbols are O for oxygen, Zn
for zinc, and Fe for iron. The first letter of a symbol is always capitalized. If the symbol contains two letters, the second letter is
lower case. The majority of elements have symbols that are based on their English names. However, some of the elements that have
been known since ancient times have maintained symbols that are based on their Latin names, as shown in Table 2.4.2.
Table 2.4.2 : Symbols and Latin Names for Elements
Chemical Symbol Name Latin Name

Na Sodium Natrium

K Potassium Kalium

Fe Iron Ferrum

Cu Copper Cuprum

Ag Silver Argentum

2.4.2 https://chem.libretexts.org/@go/page/423590
Chemical Symbol Name Latin Name

Sn Tin Stannum

Sb Antimony Stibium

Au Gold Aurum

Pb Lead Plumbum

Summary
Elements are pure substances that make up all matter, so each one is given a unique name.
The names of elements are also represented by unique one- or two-letter symbols.
Each element has a unique number of protons. An element's atomic number is equal to the number of protons in the nuclei of
any of its atoms.
The mass number of an atom is the sum of the protons and neutrons in the atom.
Isotopes are atoms of the same element (same number of protons) that have different numbers of neutrons in their atomic
nuclei.

2.4: Elements- Defined by Their Number of Protons is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-
Agnew & Henry Agnew.
4.5: Elements- Defined by Their Number of Protons by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

2.4.3 https://chem.libretexts.org/@go/page/423590
2.5: Elements- Defined by Their Number of Protons
 Learning Objectives
Define atomic number.
Define mass number.
Determine the number of protons, neutrons, and electrons in an atom.

It's important to be able to distinguish atoms of one element from atoms of another element. Elements are pure substances that
make up all other matter, so each one is given a unique name. The names of elements are also represented by unique one- or two-
letter symbols, such as H for hydrogen, C for carbon, or He for helium. However, it would more powerful if these names could be
used to identify the numbers of protons and neutrons in the atoms. That's where atomic number and mass number are useful.

Figure 2.5.1 : It is difficult to find qualities that differ between each element, and to distinguish one element from another. Each
element, however, does have a unique number of protons. Sulfur has 16 protons, silicon has 14 protons, and gold has 79 protons.
Images used with permission (public domain for sulfur and silicon, gold is licensed by CC-BY-SA-NC-ND; Alchemist-hp).

Atomic Number
Scientists distinguish between different elements by counting the number of protons in the nucleus (Table 2.5.1). If an atom has
only one proton, we know that it's a hydrogen atom. An atom with two protons is always a helium atom. If scientists count four
protons in an atom, they know it's a beryllium atom. An atom with three protons is a lithium atom, an atom with five protons is a
boron atom, an atom with six protons is a carbon atom . . . the list goes on.
Since an atom of one element can be distinguished from an atom of another element by the number of protons in its nucleus,
scientists are always interested in this number, and how this number differs between different elements. The number of protons in
an atom is called its atomic number (Z ). This number is very important because it is unique for atoms of a given element. All
atoms of an element have the same number of protons, and every element has a different number of protons in its atoms. For
example, all helium atoms have two protons, and no other elements have atoms with two protons.
Table 2.5.1 : Atoms of the First Six Elements
Name Protons Neutrons Electrons Atomic Number (Z) Mass Number (A)

Hydrogen 1 0 1 1 1

Helium 2 2 2 2 4

Lithium 3 4 3 3 7

Beryllium 4 5 4 4 9

Boron 5 6 5 5 11

Carbon 6 6 6 6 12

Of course, since neutral atoms have to have one electron for every proton, an element's atomic number also tells you how many
electrons are in a neutral atom of that element. For example, hydrogen has an atomic number of 1. This means that an atom of
hydrogen has one proton, and, if it's neutral, one electron as well. Gold, on the other hand, has an atomic number of 79, which
means that an atom of gold has 79 protons, and, if it's neutral, 79 electrons as well.

2.5.1 https://chem.libretexts.org/@go/page/423591
 Neutral Atoms

Atoms are neutral in electrical charge because they have the same number of negative electrons as positive protons (Table
2.5.1). Therefore, the atomic number of an atom also tells you how many electrons the atom has. This, in turn, determines

many of the atom's chemical properties.

Mass Number
The mass number (A ) of an atom is the total number of protons and neutrons in its nucleus. The mass of the atom is a unit called
the atomic mass unit (amu). One atomic mass unit is the mass of a proton, or about 1.67 × 10 −27
kilograms, which is an
extremely small mass. A neutron has just a tiny bit more mass than a proton, but its mass is often assumed to be one atomic mass
unit as well. Because electrons have virtually no mass, just about all the mass of an atom is in its protons and neutrons. Therefore,
the total number of protons and neutrons in an atom determines its mass in atomic mass units (Table 2.5.1).
Consider helium again. Most helium atoms have two neutrons in addition to two protons. Therefore the mass of most helium atoms
is 4 atomic mass units (2 amu for the protons + 2 amu for the neutrons). However, some helium atoms have more or less than two
neutrons. Atoms with the same number of protons but different numbers of neutrons are called isotopes. Because the number of
neutrons can vary for a given element, the mass numbers of different atoms of an element may also vary. For example, some
helium atoms have three neutrons instead of two (these are called isotopes and are discussed in detail later on).
Why do you think that the "mass number" includes protons and neutrons, but not electrons? You know that most of the mass of an
atom is concentrated in its nucleus. The mass of an atom depends on the number of protons and neutrons. You have already learned
that the mass of an electron is very, very small compared to the mass of either a proton or a neutron (like the mass of a penny
compared to the mass of a bowling ball). Counting the number of protons and neutrons tells scientists about the total mass of an
atom.

mass number A = (number of protons) + (number of neutrons)

An atom's mass number is very easy to calculate, provided that you know the number of protons and neutrons in an atom.

 Example 4.5.1

What is the mass number of an atom of helium that contains 2 neutrons?

Solution
(number of protons) = 2 (Remember that an atom of helium always has 2 protons.)
(number of neutrons) = 2

mass number = (number of protons) + (number of neutrons)

mass number = 2 + 2 = 4

A chemical symbol is a one- or two-letter designation of an element. Some examples of chemical symbols are O for oxygen, Zn
for zinc, and Fe for iron. The first letter of a symbol is always capitalized. If the symbol contains two letters, the second letter is
lower case. The majority of elements have symbols that are based on their English names. However, some of the elements that have
been known since ancient times have maintained symbols that are based on their Latin names, as shown in Table 2.5.2.
Table 2.5.2 : Symbols and Latin Names for Elements
Chemical Symbol Name Latin Name

Na Sodium Natrium

K Potassium Kalium

Fe Iron Ferrum

Cu Copper Cuprum

Ag Silver Argentum

2.5.2 https://chem.libretexts.org/@go/page/423591
Chemical Symbol Name Latin Name

Sn Tin Stannum

Sb Antimony Stibium

Au Gold Aurum

Pb Lead Plumbum

Summary
Elements are pure substances that make up all matter, so each one is given a unique name.
The names of elements are also represented by unique one- or two-letter symbols.
Each element has a unique number of protons. An element's atomic number is equal to the number of protons in the nuclei of
any of its atoms.
The mass number of an atom is the sum of the protons and neutrons in the atom.
Isotopes are atoms of the same element (same number of protons) that have different numbers of neutrons in their atomic
nuclei.

2.5: Elements- Defined by Their Number of Protons is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-
Agnew & Henry Agnew.
4.5: Elements- Defined by Their Number of Protons by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

2.5.3 https://chem.libretexts.org/@go/page/423591
2.6: Looking for Patterns - The Periodic Table
 Learning Objectives
Explain how elements are organized into the periodic table.
Describe how some characteristics of elements relate to their positions on the periodic table.

In the 19th century, many previously unknown elements were discovered, and scientists noted that certain sets of elements had
similar chemical properties. For example, chlorine, bromine, and iodine react with other elements (such as sodium) to make similar
compounds. Likewise, lithium, sodium, and potassium react with other elements (such as oxygen) to make similar compounds.
Why is this so?
In 1864, Julius Lothar Meyer, a German chemist, organized the elements by atomic mass and grouped them according to their
chemical properties. Later that decade, Dmitri Mendeleev, a Russian chemist, organized all the known elements according to
similar properties. He left gaps in his table for what he thought were undiscovered elements, and he made some bold predictions
regarding the properties of those undiscovered elements. When elements were later discovered whose properties closely matched
Mendeleev’s predictions, his version of the table gained favor in the scientific community. Because certain properties of the
elements repeat on a regular basis throughout the table (that is, they are periodic), it became known as the periodic table.

Mendeleev had to list some elements out of the order of their atomic masses to group them
with other elements that had similar properties.
The periodic table is one of the cornerstones of chemistry because it organizes all of the known elements on the basis of their
chemical properties. A modern version is shown in Figure 2.6.1. Most periodic tables provide additional data (such as atomic
mass) in a box that contains each element’s symbol. The elements are listed in order of atomic number.

Figure 2.6.1 : A Modern Periodic Table. A modern periodic table lists elements left to right by atomic number. An interactive
Periodic table can be found Periodic Table of Elements. (Public Domain; PubChem via NIH)

2.6.1 https://chem.libretexts.org/@go/page/423592
Features of the Periodic Table
Elements that have similar chemical properties are grouped in columns called groups (or families). As well as being numbered,
some of these groups have names—for example, alkali metals (the first column of elements), alkaline earth metals (the second
column of elements), halogens (the next-to-last column of elements), and noble gases (the last column of elements).

The word halogen comes from the Greek for “salt maker” because these elements
combine with other elements to form a group of compounds called salts.

 To Your Health: Radon

Radon is an invisible, odorless noble gas that is slowly released from the ground, particularly from rocks and soils whose
uranium content is high. Because it is a noble gas, radon is not chemically reactive. Unfortunately, it is radioactive, and
increased exposure to it has been correlated with an increased lung cancer risk.
Because radon comes from the ground, we cannot avoid it entirely. Moreover, because it is denser than air, radon tends to
accumulate in basements, which if improperly ventilated can be hazardous to a building’s inhabitants. Fortunately, specialized
ventilation minimizes the amount of radon that might collect. Special fan-and-vent systems are available that draw air from
below the basement floor, before it can enter the living space, and vent it above the roof of a house.
After smoking, radon is thought to be the second-biggest preventable cause of lung cancer in the United States. The American
Cancer Society estimates that 10% of all lung cancers are related to radon exposure. There is uncertainty regarding what levels
of exposure cause cancer, as well as what the exact causal agent might be (either radon or one of its breakdown products, many
of which are also radioactive and, unlike radon, not gases). The US Environmental Protection Agency recommends testing
every floor below the third floor for radon levels to guard against long-term health effects.

Each row of elements on the periodic table is called a period. Periods have different lengths; the first period has only 2 elements
(hydrogen and helium), while the second and third periods have 8 elements each. The fourth and fifth periods have 18 elements
each, and later periods are so long that a segment from each is removed and placed beneath the main body of the table.
Certain elemental properties become apparent in a survey of the periodic table as a whole. Every element can be classified as either
a metal, a nonmetal, or a metalloid (or semi metal), as shown in Figure 2.6.2. A metal is a substance that is shiny, typically (but not
always) silvery in color, and an excellent conductor of electricity and heat. Metals are also malleable (they can be beaten into thin
sheets) and ductile (they can be drawn into thin wires). A nonmetal is typically dull and a poor conductor of electricity and heat.
Solid nonmetals are also very brittle. As shown in Figure 2.6.2, metals occupy the left three-fourths of the periodic table, while
nonmetals (except for hydrogen) are clustered in the upper right-hand corner of the periodic table. The elements with properties
intermediate between those of metals and nonmetals are called metalloids (or semi-metals). Elements adjacent to the bold line in
the right-hand portion of the periodic table have semimetal properties.

Figure 2.6.2 : Types of Elements. Elements are either metals, nonmetals, or metalloids (or semi metals). Each group is located in a
different part of the periodic table. (CC BY-NC-SA; Anonymous by request)

 Example 2.6.1
Based on its position in the periodic table, classify each element below as metal, a nonmetal, or a metalloid.
a. Se

2.6.2 https://chem.libretexts.org/@go/page/423592
b. Mg
c. Ge

Solution
a. In Figure 2.6.1, selenium lies above and to the right of the diagonal line marking the boundary between metals and
nonmetals, so it should be a nonmetal.
b. Magnesium lies to the left of the diagonal line marking the boundary between metals and nonmetals, so it should be a
metal.
c. Germanium lies within the diagonal line marking the boundary between metals and nonmetals, so it should be a metalloid.

 Exercise 2.6.1

Based on its location in the periodic table, do you expect indium to be a nonmetal, a metal, or a metalloid?

Answer
Indium is a metal.

Another way to categorize the elements of the periodic table is shown in Figure 2.6.3. The first two columns on the left and the last
six columns on the right are called the main group elements. The ten-column block between these columns contains the transition
metals. The two rows beneath the main body of the periodic table contain the inner transition metals. The elements in these two
rows are also referred to as, respectively, the lanthanide metals and the actinide metals.

Figure 2.6.3 : Special Names for Sections of the Periodic Table. (CC BY-NC-SA; Anonymous by request)
Sections of period table: Alkali metals, alkaline earth metals, transition metals, halogens, noble gases, lanthanide metals, actinide
metals.

Descriptive Names
As previously noted, the periodic table is arranged so that elements with similar chemical behaviors are in the same group.
Chemists often make general statements about the properties of the elements in a group using descriptive names with historical
origins.

Group 1: The Alkali Metals


The alkali metals are lithium, sodium, potassium, rubidium, cesium, and francium. Hydrogen is unique in that it is generally placed
in Group 1, but it is not a metal.
The compounds of the alkali metals are common in nature and daily life. One example is table salt (sodium chloride); lithium
compounds are used in greases, in batteries, and as drugs to treat patients who exhibit manic-depressive, or bipolar, behavior.
Although lithium, rubidium, and cesium are relatively rare in nature, and francium is so unstable and highly radioactive that it
exists in only trace amounts, sodium and potassium are the seventh and eighth most abundant elements in Earth’s crust,
respectively.

2.6.3 https://chem.libretexts.org/@go/page/423592
Alkali metals in water - Chemical eleme…
eleme…

Video: Alkali metals in water - Chemical elements: properties and reactions. (The Open University via https://youtu.be/6ZY6d6jrq-
0)

Group 2: The Alkaline Earth Metals


The alkaline earth metals are beryllium, magnesium, calcium, strontium, barium, and radium. Beryllium, strontium, and barium are
rare, and radium is unstable and highly radioactive. In contrast, calcium and magnesium are the fifth and sixth most abundant
elements on Earth, respectively; they are found in huge deposits of limestone and other minerals.

Group 17: The Halogens


The halogens are fluorine, chlorine, bromine, iodine, and astatine. The name halogen is derived from the Greek words for “salt
forming,” which reflects that all of the halogens react readily with metals to form compounds, such as sodium chloride and calcium
chloride (used in some areas as road salt).
Compounds that contain the fluoride ion are added to toothpaste and the water supply to prevent dental cavities. Fluorine is also
found in Teflon coatings on kitchen utensils. Although chlorofluorocarbon propellants and refrigerants are believed to lead to the
depletion of Earth’s ozone layer and contain both fluorine and chlorine, the latter is responsible for the adverse effect on the ozone
layer. Bromine and iodine are less abundant than chlorine, and astatine is so radioactive that it exists in only negligible amounts in
nature.

Group 18: The Noble Gases


The noble gases are helium, neon, argon, krypton, xenon, and radon. Because the noble gases are composed of only single atoms,
they are called monatomic. At room temperature and pressure, they are unreactive gases. Because of their lack of reactivity, for
many years they were called inert gases or rare gases. However, the first chemical compounds containing the noble gases were
prepared in 1962. Although the noble gases are relatively minor constituents of the atmosphere, natural gas contains substantial
amounts of helium. Because of its low reactivity, argon is often used as an unreactive (inert) atmosphere for welding and in light
bulbs. The red light emitted by neon in a gas discharge tube is used in neon lights.

 Example 2.6.2: Groups

Provide the family or group name of each element.


a. Li
b. Ar
c. Cl
Solution
a. Lithium is an alkali metal (Group 1)
b. Argon is a noble gas (Group 18)
c. Chlorine is a halogen (Group 17)

2.6.4 https://chem.libretexts.org/@go/page/423592
 Exercise 2.6.2: Groups

Provide the family or group name of each element.


a. F
b. Ca
c. Kr

Answer a:
Fluorine is a halogen (Group 17).
Answer b:
Calcium is a alkaline earth metal (Group 2).
Answer c:
Krypton is a noble gas (Group 18).

 Example 2.6.3: Classification of Elements

Classify each element as metal, non metal, transition metal or inner transition metal.
a. Li
b. Ar
c. Am
d. Fe
Solution
a. Lithium is a metal.
b. Argon is a non metal.
c. Americium is an inner transition metal.
d. Iron is a transition metal.

 Exercise 2.6.3: Classification of Elements

Classify each element as metal, non metal, transition metal or inner transition metal.
a. F
b. U
c. Cu

Answer a:
Fluorine is a nonmetal.
Answer b:
Uranium is a metal (and a inner transition metal too).
Answer c:
Copper is a metal (and a transition metal too).

Summary
The periodic table is an arrangement of the elements in order of increasing atomic number. Elements that exhibit similar chemistry
appear in vertical columns called groups (numbered 1–18 from left to right); the seven horizontal rows are called periods. Some of
the groups have widely-used common names, including the alkali metals (Group 1) and the alkaline earth metals (Group 2) on the
far left, and the halogens (Group 17) and the noble gases (Group 18) on the far right.
The elements can be broadly divided into metals, nonmetals, and semi metals. Semi metals exhibit properties intermediate between
those of metals and nonmetals. Metals are located on the left of the periodic table, and nonmetals are located on the upper right.
They are separated by a diagonal band of semi metals. Metals are lustrous, good conductors of electricity, and readily shaped (they

2.6.5 https://chem.libretexts.org/@go/page/423592
are ductile and malleable). Solid nonmetals are generally brittle and poor electrical conductors. Other important groupings of
elements in the periodic table are the main group elements, the transition metals, and the inner transition metals (the lanthanides,
and the actinides).

References
1. Petrucci, Ralph H., William S. Harwood, F. G. Herring, and Jeffrey D. Madura. General Chemistry: Principles and Modern
Applications. 9th ed. Upper Saddle River: Pearson Education, Inc., 2007.
2. Sisler, Harry H. Electronic structure, properties, and the periodic law. New york; Reinhold publishing corporation, 1963.
3. Petrucci, Ralph H., Carey Bissonnette, F. G. Herring, and Jeffrey D. Madura. General Chemistry: Principles and Modern
Applications. Custom Edition for CHEM 2. Pearson Learning Solutions, 2010.

Contributions & Attributions

2.6: Looking for Patterns - The Periodic Table is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-
Agnew & Henry Agnew.
4.6: Looking for Patterns - The Periodic Table by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

2.6.6 https://chem.libretexts.org/@go/page/423592
2.7: Ions - Losing and Gaining Electrons
 Learning Objectives
Define the two types of ions.

Most atoms do not have eight electrons in their valence electron shell. Some atoms have only a few electrons in their outer shell,
while some atoms lack only one or two electrons to have an octet. In cases where an atom has three or fewer valence electrons, the
atom may lose those valence electrons quite easily until what remains is a lower shell that contains an octet. Atoms that lose
electrons acquire a positive charge as a result because they are left with fewer negatively charged electrons to balance the positive
charges of the protons in the nucleus. Positively charged ions are called cations. Most metals become cations when they make ionic
compounds.

Cations
A neutral sodium atom is likely to achieve an octet in its outermost shell by losing its one valence electron.
+ −
Na → Na +e

The cation produced in this way, Na+, is called the sodium ion to distinguish it from the element. The outermost shell of the sodium
ion is the second electron shell, which has eight electrons in it. The octet rule has been satisfied. Figure 2.7.1 is a graphical
depiction of this process.

Figure 2.7.1 : The Formation of a Sodium Ion. On the left, a sodium atom has 11 electrons. On the right, the sodium ion only has 10
electrons and a 1+ charge.
Neutral sodium atom on left has 11 protons and 11 electrons. Sodium ion on right has 11 protons and 10 electrons, with a +1
overall charge.

Anions
Some atoms have nearly eight electrons in their valence shell and can gain additional valence electrons until they have an octet.
When these atoms gain electrons, they acquire a negative charge because they now possess more electrons than protons. Negatively
charged ions are called anions. Most nonmetals become anions when they make ionic compounds.
A neutral chlorine atom has seven electrons in its outermost shell. Only one more electron is needed to achieve an octet in
chlorine’s valence shell. (In table salt, this electron comes from the sodium atom.)
− −
e + Cl ⟶ Cl

In this case, the ion has the same outermost shell as the original atom, but now that shell has eight electrons in it. Once again, the
octet rule has been satisfied. The resulting anion, Cl−, is called the chloride ion; note the slight change in the suffix (-ide instead of -
ine) to create the name of this anion. Figure 2.7.2 is a graphical depiction of this process.

2.7.1 https://chem.libretexts.org/@go/page/423593
Figure 2.7.2 : The Formation of a Chlorine Ion. On the left, the chlorine atom has 17 electrons. On the right, the chloride ion has 18
electrons and has a 1− charge.
Neutral chlorine atom on left has 17 protons and 17 electrons. Sodium ion on right has 17 protons and 18 electrons, with a -1
overall charge.

The names for positive and negative ions are pronounced CAT-eye-ons and ANN-eye-ons,
respectively.
In many cases, elements that belong to the same group (vertical column) on the periodic table form ions with the same charge
because they have the same number of valence electrons. Thus, the periodic table becomes a tool for remembering the charges on
many ions. For example, all ions made from alkali metals, the first column on the periodic table, have a 1+ charge. Ions made from
alkaline earth metals, the second group on the periodic table, have a 2+ charge. On the other side of the periodic table, the next-to-
last column, the halogens, form ions having a 1− charge. Figure 2.7.3 shows how the charge on many ions can be predicted by the
location of an element on the periodic table. Note the convention of first writing the number and then the sign on a ion with
multiple charges. The barium cation is written Ba2+, not Ba+2.

Figure 2.7.3 : Predicting Ionic Charges. The charge that an atom acquires when it becomes an ion is related to the structure of the
periodic table. Within a group (family) of elements, atoms form ions of a certain charge.

Contributions & Attributions

2.7: Ions - Losing and Gaining Electrons is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew &
Henry Agnew.
4.7: Ions - Losing and Gaining Electrons by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

2.7.2 https://chem.libretexts.org/@go/page/423593
2.8: Isotopes - When the Number of Neutrons Varies
 Learning Objectives
Explain what isotopes are and how an isotope affects an element's atomic mass.
Determine the number of protons, electrons, and neutrons of an element with a given mass number.

All atoms of the same element have the same number of protons, but some may have different numbers of neutrons. For example,
all carbon atoms have six protons, and most have six neutrons as well. But some carbon atoms have seven or eight neutrons instead
of the usual six. Atoms of the same element that differ in their numbers of neutrons are called isotopes. Many isotopes occur
naturally. Usually one or two isotopes of an element are the most stable and common. Different isotopes of an element generally
have the same physical and chemical properties because they have the same numbers of protons and electrons.

An Example: Hydrogen Isotopes


Hydrogen is an example of an element that has isotopes. Three isotopes of hydrogen are modeled in Figure 2.8.1. Most hydrogen
atoms have just one proton, one electron, and lack a neutron. These atoms are just called hydrogen. Some hydrogen atoms have one
neutron as well. These atoms are the isotope named deuterium. Other hydrogen atoms have two neutrons. These atoms are the
isotope named tritium.

Figure 2.8.1 : The three most stable isotopes of hydrogen: protium (A = 1), deuterium (A = 2), and tritium (A = 3). (CC SA-BY 3.0;
Balajijagadesh via Wikipedia).
For most elements other than hydrogen, isotopes are named for their mass number. For example, carbon atoms with the usual 6
neutrons have a mass number of 12 (6 protons + 6 neutrons = 12), so they are called carbon-12. Carbon atoms with 7 neutrons have
an atomic mass of 13 (6 protons + 7 neutrons = 13). These atoms are the isotope called carbon-13.

 Example 2.8.1: Lithium Isotopes


a. What is the atomic number and the mass number of an isotope of lithium containing 3 neutrons?
b. What is the atomic number and the mass number of an isotope of lithium containing 4 neutrons?

Solution
A lithium atom contains 3 protons in its nucleus irrespective of the number of neutrons or electrons.
a.
atomic number = (number of protons) = 3

(number of neutrons) = 3

mass number = (number of protons) + (number of neutrons)

mass number = 3 + 3

=6

b.

atomic number = (number of protons) = 3

(number of neutrons) = 4

2.8.1 https://chem.libretexts.org/@go/page/423594
mass number = (number of protons) + (number of neutrons)

mass number = 3 + 4

=7

Notice that because the lithium atom always has 3 protons, the atomic number for lithium is always 3. The mass number,
however, is 6 in the isotope with 3 neutrons, and 7 in the isotope with 4 neutrons. In nature, only certain isotopes exist. For
instance, lithium exists as an isotope with 3 neutrons, and as an isotope with 4 neutrons, but it doesn't exist as an isotope with 2
neutrons or as an isotope with 5 neutrons.

Stability of Isotopes
Atoms need a certain ratio of neutrons to protons to have a stable nucleus. Having too many or too few neutrons relative to protons
results in an unstable, or radioactive, nucleus that will sooner or later break down to a more stable form. This process is called
radioactive decay. Many isotopes have radioactive nuclei, and these isotopes are referred to as radioisotopes. When they decay,
they release particles that may be harmful. This is why radioactive isotopes are dangerous and why working with them requires
special suits for protection. The isotope of carbon known as carbon-14 is an example of a radioisotope. In contrast, the carbon
isotopes called carbon-12 and carbon-13 are stable.
This whole discussion of isotopes brings us back to Dalton's Atomic Theory. According to Dalton, atoms of a given element are
identical. But if atoms of a given element can have different numbers of neutrons, then they can have different masses as well!
How did Dalton miss this? It turns out that elements found in nature exist as constant uniform mixtures of their naturally occurring
isotopes. In other words, a piece of lithium always contains both types of naturally occurring lithium (the type with 3 neutrons and
the type with 4 neutrons). Moreover, it always contains the two in the same relative amounts (or "relative abundance"). In a chunk
of lithium, 93% will always be lithium with 4 neutrons, while the remaining 7% will always be lithium with 3 neutrons.
Dalton always experimented with large chunks of an element—chunks that contained all of the naturally occurring isotopes of that
element. As a result, when he performed his measurements, he was actually observing the averaged properties of all the different
isotopes in the sample. For most of our purposes in chemistry, we will do the same thing and deal with the average mass of the
atoms. Luckily, aside from having different masses, most other properties of different isotopes are similar.
There are two main ways in which scientists frequently show the mass number of an atom they are interested in. It is important to
note that the mass number is not given on the periodic table. These two ways include writing a nuclear symbol or by giving the
name of the element with the mass number written.
To write a nuclear symbol, the mass number is placed at the upper left (superscript) of the chemical symbol and the atomic
number is placed at the lower left (subscript) of the symbol. The complete nuclear symbol for helium-4 is drawn below:

Nuclear symbol for helium-4: The element symbol is He, the mass number to the top left is 4, and the atomic number to the bottom
left is 2
The following nuclear symbols are for a nickel nucleus with 31 neutrons and a uranium nucleus with 146 neutrons.
59
Ni
28

238
92
U

In the nickel nucleus represented above, the atomic number 28 indicates that the nucleus contains 28 protons, and therefore, it must
contain 31 neutrons in order to have a mass number of 59. The uranium nucleus has 92 protons, as all uranium nuclei do; and this
particular uranium nucleus has 146 neutrons.
Another way of representing isotopes is by adding a hyphen and the mass number to the chemical name or symbol. Thus the two
nuclei would be Nickel-59 or Ni-59 and Uranium-238 or U-238, where 59 and 238 are the mass numbers of the two atoms,
respectively. Note that the mass numbers (not the number of neutrons) are given to the side of the name.

2.8.2 https://chem.libretexts.org/@go/page/423594
 Example 2.8.2: Potassium-40

How many protons, electrons, and neutrons are in an atom of 40


19
?
K

Solution
atomic number = (number of protons) = 19

For all atoms with no charge, the number of electrons is equal to the number of protons.

number of electrons = 19

The mass number, 40, is the sum of the protons and the neutrons.
To find the number of neutrons, subtract the number of protons from the mass number.

number of neutrons = 40 − 19 = 21.

 Example 2.8.3: Zinc-65

How many protons, electrons, and neutrons are in an atom of zinc-65?

Solution
number of protons = 30

For all atoms with no charge, the number of electrons is equal to the number of protons.

number of electrons = 30

The mass number, 65, is the sum of the protons and the neutrons.
To find the number of neutrons, subtract the number of protons from the mass number.

number of neutrons = 65 − 30 = 35

 Exercise 2.8.3

How many protons, electrons, and neutrons are in each atom?


a. Co
60
27

b. Na-24
c. Ca
45
20

d. Sr-90

Answer a:
27 protons, 27 electrons, 33 neutrons
Answer b:

2.8.3 https://chem.libretexts.org/@go/page/423594
11 protons, 11 electrons, 13 neutrons
Answer c:
20 protons, 20 electrons, 25 neutrons
Answer d:
38 protons, 38 electrons, 52 neutrons

Summary
The number of protons is always the same in atoms of the same element.
The number of neutrons can be different, even in atoms of the same element.
Atoms of the same element that contain the same number of protons, but different numbers of neutrons, are known as isotopes.
Isotopes of any given element all contain the same number of protons, so they have the same atomic number (for example, the
atomic number of helium is always 2).
Isotopes of a given element contain different numbers of neutrons, therefore, different isotopes have different mass numbers.

2.8: Isotopes - When the Number of Neutrons Varies is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa
Alviar-Agnew & Henry Agnew.
4.8: Isotopes - When the Number of Neutrons Varies by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

2.8.4 https://chem.libretexts.org/@go/page/423594
2.9: Atomic Mass - The Average Mass of an Element’s Atoms
 Learning Objectives
Explain what is meant by the atomic mass of an element.
Calculate the atomic mass of an element from the masses and relative percentages of the isotopes of the element.

In chemistry we very rarely deal with only one isotope of an element. We use a mixture of the isotopes of an element in chemical
reactions and other aspects of chemistry, because all of the isotopes of an element react in the same manner. That means that we
rarely need to worry about the mass of a specific isotope, but instead we need to know the average mass of the atoms of an element.
Using the masses of the different isotopes and how abundant each isotope is, we can find the average mass of the atoms of an
element. The atomic mass of an element is the weighted average mass of the atoms in a naturally occurring sample of the element.
Atomic mass is typically reported in atomic mass units.

Calculating Atomic Mass


You can calculate the atomic mass (or average mass) of an element provided you know the relative abundance (the fraction of an
element that is a given isotope), the element's naturally occurring isotopes, and the masses of those different isotopes. We can
calculate this by the following equation:

Atomic mass = (%1 ) (mass1 ) + (%2 ) (mass2 ) + ⋯

Look carefully to see how this equation is used in the following examples.

 Example 2.9.1: Boron Isotopes

Boron has two naturally occurring isotopes. In a sample of boron, 20% of the atoms are B -10, which is an isotope of boron
with 5 neutrons and mass of 10 amu. The other 80% of the atoms are B -11, which is an isotope of boron with 6 neutrons and a
mass of 11 amu. What is the atomic mass of boron?

Solution
Boron has two isotopes. We will use the equation:

Atomic mass = (%1 ) (mass1 ) + (%2 ) (mass2 ) + ⋯

Isotope 1: % 1 = 0.20(Write all percentages as decimals), mass 1 = 10

Isotope 2: % 2 ,
= 0.80 mass2 = 11

Substitute these into the equation, and we get:

Atomic mass = (0.20) (10) + (0.80) (11)

Atomic mass = 10.8 amu

The mass of an average boron atom, and thus boron's atomic mass, is 10.8 amu.

 Example 2.9.2: Neon Isotopes


Neon has three naturally occurring isotopes. In a sample of neon, 90.92% of the atoms are Ne-20, which is an isotope of neon
with 10 neutrons and a mass of 19.99 amu. Another 0.3% of the atoms are Ne-21, which is an isotope of neon with 11
neutrons and a mass of 20.99 amu. The final 8.85% of the atoms are Ne-22, which is an isotope of neon with 12 neutrons and
a mass of 21.99 amu. What is the atomic mass of neon?

Solution
Neon has three isotopes. We will use the equation:

2.9.1 https://chem.libretexts.org/@go/page/423595
Atomic mass = (%1 ) (mass1 ) + (%2 ) (mass2 ) + ⋯

Isotope 1: % 1 = 0.9092 (write all percentages as decimals), mass


1 = 19.99

Isotope 2: % 2 ,
= 0.003 mass2 = 20.99

Isotope 3: % 3 ,
= 0.0885 mass3 = 21.99

Substitute these into the equation, and we get:

Atomic mass = (0.9092) (19.99) + (0.003) (20.99) + (0.0885) (21.99)

Atomic mass = 20.17 amu

The mass of an average neon atom is 20.17 amu

The periodic table gives the atomic mass of each element. The atomic mass is a number that usually appears below the element's
symbol in each square. Notice that the atomic mass of boron (symbol B ) is 10.8, which is what we calculated in Example 2.9.1,
and the atomic mass of neon (symbol Ne) is 20.8, which is what we calculated in Example 2.9.2. Take time to notice that not all
periodic tables have the atomic number above the element's symbol and the mass number below it. If you are ever confused,
remember that the atomic number should always be the smaller of the two and will be a whole number, while the atomic mass
should always be the larger of the two and will be a decimal number.

 Exercise 2.9.1

Chlorine has two naturally occurring isotopes. In a sample of chlorine, 75.77% of the atoms are Cl-35, with a mass of
34.97 amu. Another 24.23% of the atoms are Cl-37, with a mass of 36.97 amu. What is the atomic mass of chlorine?

Answer
35.45 amu

Summary
An element's atomic mass is the weighted average of the masses of the isotopes of an element
An element's atomic mass can be calculated provided the relative abundance of the element's naturally occurring isotopes and
the masses of those isotopes are known.
The periodic table is a convenient way to summarize information about the different elements. In addition to the element's
symbol, most periodic tables will also contain the element's atomic number and the element's atomic mass.

2.9: Atomic Mass - The Average Mass of an Element’s Atoms is shared under a CK-12 license and was authored, remixed, and/or curated by
Marisa Alviar-Agnew & Henry Agnew.
4.9: Atomic Mass - The Average Mass of an Element’s Atoms by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

2.9.2 https://chem.libretexts.org/@go/page/423595
2.10: Light- Electromagnetic Radiation
 Learning Objectives
Define the terms wavelength and frequency with respect to wave-form energy.
State the relationship between wavelength and frequency with respect to electromagnetic radiation.

During the summer, almost everyone enjoys going to the beach. Beach-goers can swim, have picnics, and work on their tans. But if
a person gets too much sun, they can burn. A particular set of solar wavelengths are especially harmful to the skin. This portion of
the solar spectrum is known as UV B, with wavelengths of 280-320 nm. Sunscreens are effective in protecting skin against both
the immediate skin damage and the long-term possibility of skin cancer.

Waves
Waves are characterized by their repetitive motion. Imagine a toy boat riding the waves in a wave pool. As the water wave passes
under the boat, it moves up and down in a regular and repeated fashion. While the wave travels horizontally, the boat only travels
vertically up and down. The figure below shows two examples of waves.

Figure 2.10.1 : (A) A wave consists of alternation crests and troughs. The wavelength (λ) is defined as the distance between any
two consecutive identical points on the waveform. The amplitude is the height of the wave. (B) A wave with a short wavelength
(top) has a high frequency because more waves pass a given point in a certain amount of time. A wave with a longer wavelength
(bottom) has a lower frequency.
A wave cycle consists of one complete wave—starting at the zero point, going up to a wave crest, going back down to a wave
trough, and back to the zero point again. The wavelength of a wave is the distance between any two corresponding points on
adjacent waves. It is easiest to visualize the wavelength of a wave as the distance from one wave crest to the next. In an equation,
wavelength is represented by the Greek letter lambda (λ). Depending on the type of wave, wavelength can be measured in meters,
centimeters, or nanometers (1 m = 10 nm) . The frequency, represented by the Greek letter nu (ν ), is the number of waves that
9

pass a certain point in a specified amount of time. Typically, frequency is measured in units of cycles per second or waves per
second. One wave per second is also called a Hertz (Hz) and in SI units is a reciprocal second (s ). −1

Figure B above shows an important relationship between the wavelength and frequency of a wave. The top wave clearly has a
shorter wavelength than the second wave. However, if you picture yourself at a stationary point watching these waves pass by,
more waves of the first kind would pass by in a given amount of time. Thus the frequency of the first wave is greater than that of
the second wave. Wavelength and frequency are therefore inversely related. As the wavelength of a wave increases, its frequency
decreases. The equation that relates the two is:

c = λν

The variable c is the speed of light. For the relationship to hold mathematically, if the speed of light is used in m/s, the wavelength
must be in meters and the frequency in Hertz.

 Example 2.10.1: Orange Light


The color orange within the visible light spectrum has a wavelength of about 620 nm. What is the frequency of orange light?

Solution
Solutions to Example 9.2.1

2.10.1 https://chem.libretexts.org/@go/page/423596
Steps for Problem Solving Example 2.10.1

Given :
Identify the "given" information and what the problem is asking you Wavelength (λ) = 620 nm
to "find." Speed of light (c) = 3.00 × 10 8
m/s

Find: Frequency (Hz)

List other known quantities. 1 m = 10


9
nm

1.Convert the wavelength to m .


2. Apply the equation c = λν and solve for frequency. Dividing both
Identify steps to get the final answer. sides of the equation by λ yields:
c
ν =
λ

1 m
−7
620 nm × ( ) = 6.20 × 10 m
9
10 nm
Cancel units and calculate. 8
c 3.0 × 10 m/s
14
ν = = = 4.8 × 10 Hz
λ −7
6.20 × 10

Think about your result. The value for the frequency falls within the range for visible light.

 Exercise 2.10.1

What is the wavelength of light if its frequency is 1.55 × 1010 s−1?

Answer
0.0194 m, or 19.4 mm

Summary
All waves can be defined in terms of their frequency and intensity. c = λν expresses the relationship between wavelength and
frequency.

2.10: Light- Electromagnetic Radiation is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew &
Henry Agnew.
9.2: Light- Electromagnetic Radiation by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

2.10.2 https://chem.libretexts.org/@go/page/423596
2.11: THE ELECTROMAGNETIC SPECTRUM
 LEARNING OBJECTIVES
Know the properties of different types of electromagnetic radiation.

Electromagnetic waves have an extremely wide range of wavelengths, frequencies, and energies. The highest energy form of
electromagnetic waves are gamma (γ) rays and the lowest energy form are radio waves.
The figure below shows the electromagnetic spectrum, which is all forms of electromagnetic radiation. On the far left of Figure 2.11.1 are
the highest energy electromagnetic waves. These are called gamma rays and can be quite dangerous, in large numbers, to living systems.
The next lower energy form of electromagnetic waves are called x-rays. Most of you are familiar with the penetration abilities of these
waves. They can also be dangerous to living systems. Humans are advised to limit as much as possible the number of medical x-rays they
have per year. Next lower, in energy, are ultraviolet rays. These rays are part of sunlight and the upper end of the ultraviolet range can
cause sunburn and perhaps skin cancer. The tiny section next in the spectrum is the visible range of light. The visible light spectrum has
been greatly expanded in the bottom half of the figure so that it can be discussed in more detail. The visible range of electromagnetic
radiation are the frequencies to which the human eye responds. Lower in the spectrum are infrared rays and radio waves.

Figure 2.11.1: The electromagnetic spectrum, with its various regions labeled. The borders of each region are approximate. (CC BY-NC-
SA; anonymous by request).
The light energies that are in the visible range are electromagnetic waves that cause the human eye to respond when those frequencies enter
the eye. The eye sends a signal to the brain and the individual "sees" various colors. The highest energy waves in the visible region cause
the brain to see violet and as the energy decreases, the colors change to blue, green, yellow, orange, and red. When the energy of the wave is
above or below the visible range, the eye does not respond to them. When the eye receives several different frequencies at the same time,
the colors are blended by the brain. If all frequencies of light strike the eye together, the brain sees white. If there are no visible frequencies
striking the eye, the brain sees black. The objects that you see around you are light absorbers—that is, the chemicals on the surface of the
object will absorb certain frequencies and not others. Your eyes detect the frequencies that strike your eye. Therefore, if your friend is
wearing a red shirt, it means the dye in that shirt absorbs every frequency except red and the red frequencies are reflected. If your only light
source was one exact frequency of blue light and you shined it on a shirt that was red in sunlight, the shirt would appear black because no
light would be reflected. The light from fluorescent types of lights do not contain all the frequencies of sunlight and so clothes inside a store
may appear to be a slightly different color when you get them home.

SUMMARY
Electromagnetic radiation has a wide spectrum, including gamma rays, X-rays, UV rays, visible light, IR radiation, microwaves, and
radio waves.
The different colors of light differ in their frequencies (or wavelengths).

2.11.1 https://chem.libretexts.org/@go/page/423597
2.11: The Electromagnetic Spectrum is shared under a CC BY-NC-SA 3.0 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew &
Henry Agnew.
9.3: The Electromagnetic Spectrum by Henry Agnew, Marisa Alviar-Agnew is licensed CC BY-NC-SA 3.0.

2.11.2 https://chem.libretexts.org/@go/page/423597
2.12: The Bohr Model - Atoms with Orbits
 Learning Objectives
Define an energy level in terms of the Bohr model.
Discuss how the Bohr model can be used to explain atomic spectra.

Electric light bulbs contain a very thin wire in them that emits light when heated. The wire is called a filament. The particular wire
used in light bulbs is made of tungsten. A wire made of any metal would emit light under these circumstances, but tungsten was
chosen because the light it emits contains virtually every frequency and therefore, the light emitted by tungsten appears white. A
wire made of some other element would emit light of some color that was not convenient for our uses. Every element emits light
when energized by heating or passing electric current through it. Elements in solid form begin to glow when they are heated
sufficiently, and elements in gaseous form emit light when electricity passes through them. This is the source of light emitted by
neon signs and is also the source of light in a fire.

Figure 2.12.1 : Human/Need/Desire. Neon sculpture by Bruce Nauman (1983), who has been characterized as a conceptual artist.

Each Element Has a Unique Spectrum


The light frequencies emitted by atoms are mixed together by our eyes so that we see a blended color. Several physicists, including
Angstrom in 1868 and Balmer in 1875, passed the light from energized atoms through glass prisms in such a way that the light was
spread out so they could see the individual frequencies that made up the light. The emission spectrum (or atomic spectrum) of a
chemical element is the unique pattern of light obtained when the element is subjected to heat or electricity.

Figure 2.12.2 : Atomic Emission Spectrum of Hydrogen.


When hydrogen gas is placed into a tube and electric current passed through it, the color of emitted light is pink. But when the
color is spread out, we see that the hydrogen spectrum is composed of four individual frequencies. The pink color of the tube is the
result of our eyes blending the four colors. Every atom has its own characteristic spectrum; no two atomic spectra are alike. The
image below shows the emission spectrum of iron. Because each element has a unique emission spectrum, elements can be defined
using them.

Figure 2.12.3 : Atomic Emission Spectrum of Iron.


You may have heard or read about scientists discussing what elements are present in the sun or some more distant star, and after
hearing that, wondered how scientists could know what elements were present in a place no one has ever been. Scientists determine
what elements are present in distant stars by analyzing the light that comes from stars and finding the atomic spectrum of elements
in that light. If the exact four lines that compose hydrogen's atomic spectrum are present in the light emitted from the star, that
element contains hydrogen.

2.12.1 https://chem.libretexts.org/@go/page/423598
Bohr's Model of the Atom
By 1913, the concept of the atom had evolved from Dalton's indivisible spheres idea, to J. J. Thomson's plum pudding model, and
then to Rutherford's nuclear atom theory. Rutherford, in addition to carrying out the brilliant experiment that demonstrated the
presence of the atomic nucleus, also proposed that the electrons circled the nucleus in a planetary type motion. The solar system or
planetary model of the atom was attractive to scientists because it was similar to something with which they were already familiar,
namely the solar system.

Figure 2.12.3 : Niels Bohr with Albert Einstein at Paul Ehrenfest's home in Leiden (December 1925).
Unfortunately, there was a serious flaw in the planetary model. It was already known that when a charged particle (such as an
electron) moves in a curved path, it gives off some form of light and loses energy in doing so. This is, after all, how we produce TV
signals. If the electron circling the nucleus in an atom loses energy, it would necessarily have to move closer to the nucleus as it
loses energy, and would eventually crash into the nucleus. Furthermore, Rutherford's model was unable to describe how electrons
give off light forming each element's unique atomic spectrum. These difficulties cast a shadow on the planetary model and
indicated that, eventually, it would have to be replaced.
In 1913, the Danish physicist Niels Bohr proposed a model of the electron cloud of an atom in which electrons orbit the nucleus
and were able to produce atomic spectra. Understanding Bohr's model requires some knowledge of electromagnetic radiation (or
light).

Energy Levels
Bohr's key idea in his model of the atom is that electrons occupy definite orbitals that require the electron to have a specific amount
of energy. In order for an electron to be in the electron cloud of an atom, it must be in one of the allowable orbitals and it must have
the precise energy required for that orbit. Orbits closer to the nucleus would require smaller amounts of energy for an electron, and
orbits farther from the nucleus would require the electron to have a greater amount of energy. The possible orbits are known as
energy levels. One of the weaknesses of Bohr's model was that he could not offer a reason why only certain energy levels or orbits
were allowed.

2.12.2 https://chem.libretexts.org/@go/page/423598
Figure 2.12.4 : The energy levels of the electrons can be viewed as rungs on a ladder. Note that the spacing between rungs gets
smaller at higher energies (CC BY-NC; Ümit Kaya)
Bohr hypothesized that the only way electrons could gain or lose energy would be to move from one energy level to another, thus
gaining or losing precise amounts of energy. The energy levels are quantized, meaning that only specific amounts are possible. It
would be like a ladder that had rungs only at certain heights. The only way you can be on that ladder is to be on one of the rungs,
and the only way you could move up or down would be to move to one of the other rungs. Suppose we had such a ladder with 10
rungs. Other rules for the ladder are that only one person can be on a rung in the normal state, and the ladder occupants must be on
the lowest rung available. If the ladder had five people on it, they would be on the lowest five rungs. In this situation, no person
could move down because all of the lower rungs are full. Bohr worked out rules for the maximum number of electrons that could
be in each energy level in his model, and required that an atom in its normal state (ground state) had all electrons in the lowest
energy levels available. Under these circumstances, no electron could lose energy because no electron could move down to a lower
energy level. In this way, Bohr's model explained why electrons circling the nucleus did not emit energy and spiral into the nucleus.

Figure 2.12.5 : In Bohr's Model of the atom, electrons absorb energy to move to a higher level and release energy to move to lower
levels. (CC BY-SA 3.0; Kurzon).

2.12.3 https://chem.libretexts.org/@go/page/423598
Bohr's Model and Atomic Spectra
The evidence used to support Bohr's model came from the atomic spectra. He suggested that an atomic spectrum is made by the
electrons in an atom moving energy levels. The electrons typically have the lowest energy possible, called the ground state. If the
electrons are given energy (through heat, electricity, light, etc.) the electrons in an atom could absorb energy by jumping to a higher
energy level, or excited state. The electrons then give off the energy in the form of a piece of light—called a photon—that they
had absorbed, to fall back to a lower energy level. The energy emitted by electrons dropping back to lower energy levels will
always be precise amounts of energy, because the differences in energy levels are precise. This explains why you see specific lines
of light when looking at an atomic spectrum—each line of light matches a specific "step down" that an electron can take in that
atom. This also explains why each element produces a different atomic spectrum. Because each element has different acceptable
energy levels for its electrons, the possible steps each element's electrons can take differ from all other elements.

Summary
Bohr's model suggests each atom has a set of unchangeable energy levels, and electrons in the electron cloud of that atom must
be in one of those energy levels.
Bohr's model suggests that the atomic spectra of atoms is produced by electrons gaining energy from some source, jumping up
to a higher energy level, then immediately dropping back to a lower energy level and emitting the energy difference between the
two energy levels.
The existence of the atomic spectra is support for Bohr's model of the atom.
Bohr's model was only successful in calculating energy levels for the hydrogen atom.

Vocabulary
Emission spectrum (or atomic spectrum) - The unique pattern of light given off by an element when it is given energy.
Energy levels - Possible orbits that an electron can have in the electron cloud of an atom.
Ground state - To be in the lowest energy level possible.
Excited state - To be in a higher energy level.
Photon - A piece of electromagnetic radiation, or light, with a specific amount of energy.
Quantized - To have a specific amount.

2.12: The Bohr Model - Atoms with Orbits is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew
& Henry Agnew.
9.4: The Bohr Model - Atoms with Orbits by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

2.12.4 https://chem.libretexts.org/@go/page/423598
2.13: The Quantum-Mechanical Model- Atoms with Orbitals
 Learning Objectives
Define quantum mechanics
Differentiate between an orbit and an orbital.

How do you study something that seemingly makes no sense? We talk about electrons being in orbits and it sounds like we can tell
where that electron is at any moment. We can draw pictures of electrons in orbit, but the reality is that we don't know exactly where
they are. We are going to take a look at an area of science that even leaves scientists puzzled. When asked about quantum
mechanics, Niels Bohr (who proposed the Bohr model of the atom) said: "Anyone who is not shocked by quantum theory has not
understood it". Richard Feynman (one of the founders of modern quantum theory) stated: "I think I can safely say that nobody
understands quantum theory." So, let's take a short trip into a land that challenges our everyday world.

Quantum Mechanics
The study of motion of large objects such as baseballs is called mechanics, or more specifically, classical mechanics. Because of
the quantum nature of the electron and other tiny particles moving at high speeds, classical mechanics is inadequate to accurately
describe their motion. Quantum mechanics is the study of the motion of objects that are atomic or subatomic in size and thus
demonstrate wave-particle duality. In classical mechanics, the size and mass of the objects involved effectively obscures any
quantum effects, so that such objects appear to gain or lose energies in any amounts. Particles whose motion is described by
quantum mechanics gain or lose energy in small pieces called quanta.
One of the fundamental (and hardest to understand) principles of quantum mechanics is that the electron is both a particle and a
wave. In the everyday macroscopic world of things we can see, something cannot be both. But this duality can exist in the quantum
world of the submicroscopic on the atomic scale.
At the heart of quantum mechanics is the idea that we cannot accurately specify the location of an electron. All we can say is that
there is a probability that it exists within this certain volume of space. The scientist Erwin Schrödinger developed an equation that
deals with these calculations, which we will not pursue at this time.

Erwin Schrödinger.
Recall that in the Bohr model, the exact path of the electron was restricted to very well-defined circular orbits around the nucleus.
An orbital is the quantum mechanical refinement of Bohr’s orbit. In contrast to his concept of a simple circular orbit with a fixed
radius, orbitals are mathematically derived regions of space with different probabilities of having an electron.

Summary
Quantum mechanics involves the study of material at the atomic level. This field deals with probabilities, since we cannot
definitely locate a particle. Orbitals are mathematically derived regions of space with different probabilities of having an electron.

2.13: The Quantum-Mechanical Model- Atoms with Orbitals is shared under a CK-12 license and was authored, remixed, and/or curated by
Marisa Alviar-Agnew & Henry Agnew.
9.5: The Quantum-Mechanical Model- Atoms with Orbitals by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

2.13.1 https://chem.libretexts.org/@go/page/423599
2.14: Quantum-Mechanical Orbitals and Electron Configurations
 Learning Objectives
Represent the organization of electrons by an electron configuration and orbital diagram.

The flight path of a commercial airliner is carefully regulated by the Federal Aviation Administration. Each airplane must maintain
a distance of five miles from another plane flying at the same altitude and 2,000 feet above and below another aircraft (1,000 feet if
the altitude is less than 29,000 feet). So, each aircraft only has certain positions it is allowed to maintain while it flies. As we
explore quantum mechanics, we see that electrons have similar restrictions on their locations.

Orbitals
We can apply our knowledge of quantum numbers to describe the arrangement of electrons for a given atom. We do this with
something called electron configurations. They are effectively a map of the electrons for a given atom. We look at the four
quantum numbers for a given electron and then assign that electron to a specific orbital below.

s Orbitals
For any value of n , a value of l = 0 places that electron in an s orbital. This orbital is spherical in shape:

Figure 2.14.1 : s orbitals have no orientational preference and resemble spheres.

p Orbitals
For the table below, we see that we can have three possible orbitals when l = 1 . These are designated as p orbitals and have
dumbbell shapes. Each of the p orbitals has a different orientation in three-dimensional space.

Figure 2.14.2 : p orbitals have an orientational preference and resemble dumbbells.

d Orbitals
When l = 2 , m values can be
l −2, −1, 0, +1, +2 for a total of five d orbitals. Note that all five of the orbitals have specific
three-dimensional orientations.

Figure 2.14.3 : d orbitals have an orientational preference and exhibit complex structures.

f Orbitals
The most complex set of orbitals are the f orbitals. When l = 3 , m values can be −3, −2,
l −1, 0, +1, +2, +3 for a total of
seven different orbital shapes. Again, note the specific orientations of the different f orbitals.

2.14.1 https://chem.libretexts.org/@go/page/423600
Figure 2.14.4 : f orbitals have an orientational preference and exhibit quite complex structures.
Orbitals that have the same value of the principal quantum number form a shell. Orbitals within a shell are divided into subshells
that have the same value of the angular quantum number. Some of the allowed combinations of quantum numbers are compared in
Table 2.14.1.
Table 2.14.1 : Electron Arrangement Within Energy Levels
Number of Orbitals Number of Electrons
Principal Quantum Number of Orbitals Number of Electrons
Allowable Sublevels per Principal Energy per Principal Energy
Number (n) per Sublevel per Sublevel
Level Level

1 s 1 1 2 2

s 1 2
2 4 8
p 3 6

s 1 2

3 p 3 9 6 18

d 5 10

s 1 2

p 3 6
4 16 32
d 5 10

f 7 14

Electron Configurations
Can you name one thing that easily distinguishes you from the rest of the world? And we're not talking about DNA—that's a little
expensive to sequence. For many people, it is their email address. Your email address allows people all over the world to contact
you. It does not belong to anyone else, but serves to identify you. Electrons also have a unique set of identifiers in the quantum
numbers that describe their location and spin. Chemists use an electronic configuration to represent the organization of electrons
in shells and subshells in an atom. An electron configuration simply lists the shell and subshell labels, with a right superscript
giving the number of electrons in that subshell. The shells and subshells are listed in the order of filling. Electrons are typically
organized around an atom by starting at the lowest possible quantum numbers first, which are the shells-subshells with lower
energies.
For example, an H atom has a single electron in the 1s subshell. Its electron configuration is
1
H : 1s

He has two electrons in the 1s subshell. Its electron configuration is


2
He : 1s

The three electrons for Li are arranged in the 1s subshell (two electrons) and the 2s subshell (one electron). The electron
configuration of Li is
2 1
Li : 1 s 2 s

Be has four electrons, two in the 1s subshell and two in the 2s subshell. Its electron configuration is
2 2
Be : 1 s 2 s

2.14.2 https://chem.libretexts.org/@go/page/423600
Now that the 2s subshell is filled, electrons in larger atoms must go into the 2p subshell, which can hold a maximum of six
electrons. The next six elements progressively fill up the 2p subshell:
B: 1s22s22p1
C: 1s22s22p2
N: 1s22s22p3
O: 1s22s22p4
F: 1s22s22p5
Ne: 1s22s22p6
Now that the 2p subshell is filled (all possible subshells in the n = 2 shell), the next electron for the next-larger atom must go into
the n = 3 shell, s subshell.

Second Period Elements


Periods refer to the horizontal rows of the periodic table. Looking at a periodic table you will see that the first period contains only
the elements hydrogen and helium. This is because the first principal energy level consists of only the s sublevel and so only two
electrons are required in order to fill the entire principal energy level. Each time a new principal energy level begins, as with the
third element lithium, a new period is started on the periodic table. As one moves across the second period, electrons are
successively added. With beryllium (Z = 4) , the 2s sublevel is complete and the 2p sublevel begins with boron (Z = 5) . Since
there are three 2p orbitals and each orbital holds two electrons, the 2p sublevel is filled after six elements. Table 2.14.1 shows the
electron configurations of the elements in the second period.
Table 2.14.2 : Electron Configurations of Second-Period Elements
Element Name Symbol Atomic Number Electron Configuration

Lithium Li 3 1s 2s
2 1

Beryllium Be 4 1s 2s
2 2

Boron B 5 2
1s 2s 2p
2 1

Carbon C 6 2
1s 2s 2p
2 2

Nitrogen N 7 2
1s 2s 2p
2 3

Oxygen O 8 2
1s 2s 2p
2 4

Fluorine F 9 2
1s 2s 2p
2 5

Neon Ne 10 2
1s 2s 2p
2 6

Aufbau Principle
Construction of a building begins at the bottom. The foundation is laid and the building goes up step by step. You obviously cannot
start with the roof since there is no place to hang it. The building goes from the lowest level to the highest level in a systematic
way. In order to create ground state electron configurations for any element, it is necessary to know the way in which the atomic
sublevels are organized in order of increasing energy. Figure 2.14.5 shows the order of increasing energy of the sublevels.
The lowest energy sublevel is always the 1s sublevel, which consists of one orbital. The single electron of the hydrogen atom will
occupy the 1s orbital when the atom is in its ground state. As we proceed with atoms with multiple electrons, those electrons are
added to the next lowest sublevel: 2s, 2p, 3s, and so on. The Aufbau principle states that an electron occupies orbitals in order
from lowest energy to highest. The Aufbau (German: "building up, construction") principle is sometimes referred to as the
"building up" principle. It is worth noting that in reality atoms are not built by adding protons and electrons one at a time, and that
this method is merely an aid for us to understand the end result.

2.14.3 https://chem.libretexts.org/@go/page/423600
Figure 2.14.5 : Electrons are added to atomic orbitals in order from low energy (bottom of the graph) to high (top of the graph)
according to the Aufbau principle. Principle energy levels are color coded, while sublevels are grouped together and each circle
represents an orbital capable of holding two electrons.
As seen in the figure above, the energies of the sublevels in different principal energy levels eventually begin to overlap. After the
3p sublevel, it would seem logical that the 3d sublevel should be the next lowest in energy. However, the 4s sublevel is slightly

lower in energy than the 3d sublevel and thus fills first. Following the filling of the 3d sublevel is the 4p, then the 5s and the 4d.
Note that the 4f sublevel does not fill until just after the 6s sublevel. Figure 2.14.6 is a useful and simple aid for keeping track of
the order of fill of the atomic sublevels.

Figure 2.14.6 : The arrow leads through each subshell in the appropriate filling order for electron configurations. This chart is
straightforward to construct. Simply make a column for all the s orbitals with each n shell on a separate row. Repeat for p, d, and f.
Be sure to only include orbitals allowed by the quantum numbers (no 1p or 2d, and so forth). Finally, draw diagonal lines from top
to bottom as shown.

2.14.4 https://chem.libretexts.org/@go/page/423600
Energy levels, sublevels, & orbitals

Video 2.14.1 : Energy levels, sublevels and orbitals.

 Example 2.14.1: Nitrogen Atoms

Nitrogen has 7 electrons. Write the electron configuration for nitrogen.


Solution:
Take a close look at Figure 2.14.5, and use it to figure out how many electrons go into each sublevel, and also the order in
which the different sublevels get filled.
1. Begin by filling up the 1s sublevel. This gives 1s2. Now all of the orbitals in the red n = 1 block are filled.

Since we used 2 electrons, there are 7 − 2 = 5 electrons left

2. Next, fill the 2s sublevel. This gives 1s22s2. Now all of the orbitals in the s sublevel of the orange n = 2 block are filled.

Since we used another 2 electrons, there are 5 − 2 = 3 electrons left

3. Notice that we haven't filled the entire n = 2 block yet… there are still the p orbitals!

The final 3 electrons go into the 2p sublevel. This gives 1s22s22p3

The overall electron configuration is: 1s22s22p3.

 Example 2.14.2: Potassium Atoms

Potassium has 19 electrons. Write the electron configuration code for potassium.

Solution
This time, take a close look at Figure 2.14.5.
1. Begin by filling up the 1s sublevel. This gives 1s2. Now the n = 1 level is filled.

Since we used 2 electrons, there are 19 − 2 = 17 electrons left

2. Next, fill the 2s sublevel. This gives 1s22s2

Since we used another 2 electrons, there are 17 − 2 = 15 electrons left

3. Next, fill the 2p sublevel. This gives 1s22s22p6. Now the n = 2 level is filled.

Since we used another 6 electrons, there are 15 − 6 = 9 electrons left

2.14.5 https://chem.libretexts.org/@go/page/423600
4. Next, fill the 3s sublevel. This gives 1s22s22p63s2

Since we used another 2 electrons, there are 9 − 2 = 7 electrons left

5. Next, fill the 3p sublevel. This gives 1s22s22p63s23p6

Since we used another 6 electrons, there are 7 − 6 = 1 electron left

Here's where we have to be careful – right after 3p6!

Remember, 4s comes before 3d

6. The final electron goes into the 4s sublevel. This gives 1s22s22p63s23p64s1
The overall electron configuration is: 1s22s22p63s23p64s1

 Exercise 2.14.1: Magnesium and Sodium Atoms

What is the electron configuration for Mg and Na?

Answer Mg
Mg: 1s22s22p63s2

Answer Na
Na: 1s22s22p63s1

Pauli Exclusion Principle


When we look at the orbital possibilities for a given atom, we see that there are different arrangements of electrons for each
different type of atom. Since each electron must maintain its unique identity, we intuitively sense that the four quantum numbers
for any given electron must not match up exactly with the four quantum numbers for any other electron in that atom.
For the hydrogen atom, there is no problem since there is only one electron in the H atom. However, when we get to helium we see
that the first three quantum numbers for the two electrons are the same: same energy level, same spherical shape. What
1 1
differentiates the two helium electrons is their spin. One of the electrons has a + spin while the other electron has a − spin. So
2 2
the two electrons in the 1s orbital are each unique and distinct from one another because their spins are different. This observation
leads to the Pauli exclusion principle, which states that no two electrons in an atom can have the same set of four quantum
numbers. The energy of the electron is specified by the principal, angular momentum, and magnetic quantum numbers. If those
three numbers are identical for two electrons, the spin numbers must be different in order for the two electrons to be differentiated
from one another. The two values of the spin quantum number allow each orbital to hold two electrons. Figure 2.14.7 shows how
the electrons are indicated in a diagram.

Figure 2.14.7 : In an orbital filling diagram, a square represents an orbital, while arrows represent electrons. An arrow pointing
upward represents one spin direction, while an arrow pointing downward represents the other spin direction.

Hund's Rule
The last of the three rules for constructing electron arrangements requires electrons to be placed one at a time in a set of orbitals
within the same sublevel. This minimizes the natural repulsive forces that one electron has for another. Hund's rule states that
orbitals of equal energy are each occupied by one electron before any orbital is occupied by a second electron and that each of the
single electrons must have the same spin. The figure below shows how a set of three p orbitals is filled with one, two, three, and
four electrons.

2.14.6 https://chem.libretexts.org/@go/page/423600
Figure 2.14.8 : The 2p sublevel, for the elements boron (Z = 5), carbon (Z = 6), nitrogen (Z = 7), and oxygen (Z = 8).
According to Hund's rule, as electrons are added to a set of orbitals of equal energy, one electron enters each orbital before any
orbital receives a second electron.

Orbital Filling Diagrams


An orbital filling diagram is the more visual way to represent the arrangement of all the electrons in a particular atom. In an
orbital filling diagram, the individual orbitals are shown as circles (or squares) and orbitals within a sublevel are drawn next to each
other horizontally. Each sublevel is labeled by its principal energy level and sublevel. Electrons are indicated by arrows inside of
the circles. An arrow pointing upwards indicates one spin direction, while a downward pointing arrow indicates the other direction.
The orbital filling diagrams for hydrogen, helium, and lithium are shown in the figure below.

Figure 2.14.9 : Orbital filling diagrams for hydrogen, helium, and lithium.
According to the Aufbau process, sublevels and orbitals are filled with electrons in order of increasing energy. Since the s sublevel
consists of just one orbital, the second electron simply pairs up with the first electron as in helium. The next element is lithium and
necessitates the use of the next available sublevel, the 2s.
The filling diagram for carbon is shown in Figure 2.14.10 . There are two 2p electrons for carbon and each occupies its own 2p

orbital.

Figure 2.14.10: Orbital filling diagram for carbon.


Oxygen has four 2p electrons. After each 2p orbital has one electron in it, the fourth electron can be placed in the first 2p orbital
with a spin opposite that of the other electron in that orbital.

Figure 2.14.11: Orbital filling diagram for oxygen.


If you keep your papers in manila folders, you can pick up a folder and see how much it weighs. If you want to know how many
different papers (articles, bank records, or whatever else you keep in a folder), you have to take everything out and count. A
computer directory, on the other hand, tells you exactly how much you have in each file. We can get the same information on
atoms. If we use an orbital filling diagram, we have to count arrows. When we look at electron configuration data, we simply add
up the numbers.

2.14.7 https://chem.libretexts.org/@go/page/423600
 Example 2.14.3: Carbon Atoms
Draw the orbital filling diagram for carbon and write its electron configuration.

Solution

Step 1: List the known quantities and plan the problem.

Known
Atomic number of carbon, Z=6
Use the order of fill diagram to draw an orbital filling diagram with a total of six electrons. Follow Hund's rule. Write the
electron configuration.

Step 2: Construct the diagram.

Orbital filling diagram for carbon.


Electron configuration 1s22s22p2

Step 3: Think about your result.


Following the 2s sublevel is the 2p, and p sublevels always consist of three orbitals. All three orbitals need to be drawn
even if one or more is unoccupied. According to Hund's rule, the sixth electron enters the second of those p orbitals and
has the same spin as the fifth electron.

 Exercise 2.14.2: Electronic Configurations

Write the electron configurations and orbital diagrams for


a. Potassium atom: K
b. Arsenic atom: As
c. Phosphorus atom: P

Answer a:
Potassium: 1s 2 2 6
2s 2p 3s 3p 4s
2 6 1

Answer b:
Arsenic: 1s 2 2 6
2s 2p 3s 3p 4s 3d
2 6 2 10 3
4p

2.14.8 https://chem.libretexts.org/@go/page/423600
Answer c:
Phosphorus 1s 2 2 6
2s 2p 3s 3p
2 3

 The Atom Neighborhood

Figure 2.14.12: The atom neighborhood. Source: Dr. Binh Dao, Sacramento City College.

Summary
There are four different classes of electron orbitals. These orbitals are determined by the value of the angular momentum quantum
number ℓ. An orbital is a wave function for an electron defined by the three quantum numbers, n, ℓ and m ℓ . Orbitals define regions
in space where you are likely to find electrons. s orbitals (ℓ = 0) are spherical shaped. p orbitals (ℓ = 1) are dumb-bell shaped. The
three possible p orbitals are always perpendicular to each other.
Electron configuration notation simplifies the indication of where electrons are located in a specific atom. Superscripts are used to
indicate the number of electrons in a given sublevel. The Aufbau principle gives the order of electron filling in an atom. It can be
used to describe the locations and energy levels of every electron in a given atom. Hund's rule specifies the order of electron filling
within a set of orbitals. Orbital filling diagrams are a way of indicating electron locations in orbitals. The Pauli exclusion principle
specifies limits on how identical quantum numbers can be for two electrons in the same atom.

Vocabulary
principal quantum number (n)

2.14.9 https://chem.libretexts.org/@go/page/423600
Defines the energy level of the wave function for an electron, the size of the electron's standing wave, and the number of nodes
in that wave.
quantum numbers
Integer numbers assigned to certain quantities in the electron wave function. Because electron standing waves must be
continuous and must not "double over" on themselves, quantum numbers are restricted to integer values.

Contributions & 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-
2bd...a7ac8df6@9.110).

2.14: Quantum-Mechanical Orbitals and Electron Configurations is shared under a CK-12 license and was authored, remixed, and/or curated by
Marisa Alviar-Agnew & Henry Agnew.
9.6: Quantum-Mechanical Orbitals and Electron Configurations by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original
source: https://www.ck12.org/c/chemistry/.

2.14.10 https://chem.libretexts.org/@go/page/423600
2.15: The Explanatory Power of the Quantum-Mechanical Model

 Learning Objectives
Give the name and location of specific groups on the periodic table, including alkali metals, alkaline earth metals, noble
gases, halogens, and transition metals.
Explain the relationship between the chemical behavior of families in the periodic table and their electron configurations.
Identify elements that will have the most similar properties to a given element.

The chemical behavior of atoms is controlled by their electron configuration. Since the families of elements were organized by
their chemical behavior, it is predictable that the individual members of each chemical family will have similar electron
configurations.

Families of the Periodic Table


Remember that Mendeleev arranged the periodic table so that elements with the most similar properties were placed in the same
group. A group is a vertical column of the periodic table. All of the 1A elements have one valence electron. This is what causes
these elements to react in the same ways as the other members of the family. The elements in 1A are all very reactive and form
compounds in the same ratios with similar properties with other elements. Because of their similarities in their chemical properties,
Mendeleev put these elements into the same group. Group 1A is also known as the alkali metals. Although most metals tend to be
very hard, these metals are actually soft and can be easily cut.
Group 2A is also called the alkaline earth metals. Once again, because of their similarities in electron configurations, these
elements have similar properties to each other. The same pattern is true of other groups on the periodic table. Remember,
Mendeleev arranged the table so that elements with the most similar properties were in the same group on the periodic table.
It is important to recognize a couple of other important groups on the periodic table by their group name. Group 7A (or 17)
elements are also called halogens. This group contains very reactive nonmetal elements.
The noble gases are in group 8A. These elements also have similar properties to each other, the most significant property being that
they are extremely unreactive, rarely forming compounds. The reason for this will be communicated later, when we discuss how
compounds form. The elements in this group are also gases at room temperature.

An alternate numbering system numbers all of the s , p, and d block elements from 1-18. In this numbering system, group 1A is
group 1; group 2A is group 2; the halogens (7A) are group 17; and the noble gases (8A) are group 18. You will come across
periodic tables with both numbering systems. It is important to recognize which numbering system is being used, and to be able to
find the number of valence electrons in the main block elements, regardless of which numbering system is being used.

Periods of the Periodic Table


If you can locate an element on the Periodic Table, you can use the element's position to
figure out the energy level of the element's valence electrons. A period is a horizontal row of
elements on the periodic table. For example, the elements sodium (Na) and magnesium (
Mg ) are both in period 3. The elements astatine (At ) and radon (Rn) are both in period 6.

2.15.1 https://chem.libretexts.org/@go/page/423601
Summary
The vertical columns on the periodic table are called groups or families because of their similar chemical behavior.
All the members of a family of elements have the same number of valence electrons and similar chemical properties.
The horizontal rows on the periodic table are called periods.

Vocabulary
Group (family) - A vertical column of the periodic table.
Alkali metals - Group 1A of the periodic table.
Alkaline earth metals - Group 2A of the periodic table.
Halogens - Group 7A of the periodic table.
Noble gases - Group 8A of the periodic table.
Transition elements - Groups 3 to 12 of the periodic table.

2.15: The Explanatory Power of the Quantum-Mechanical Model is shared under a CK-12 license and was authored, remixed, and/or curated by
Marisa Alviar-Agnew & Henry Agnew.
9.8: The Explanatory Power of the Quantum-Mechanical Model by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original
source: https://www.ck12.org/c/chemistry/.

2.15.2 https://chem.libretexts.org/@go/page/423601
2.16: Periodic Trends - Atomic Size, Ionization Energy, and Metallic Character
 Learning Objectives
Be able to state how certain properties of atoms vary based on their relative position on the periodic table.

One of the reasons the periodic table is so useful is because its structure allows us to qualitatively determine how some properties
of the elements vary versus their position on the periodic table. The variations of properties versus positions on the periodic table
are called periodic trends. There is no other tool in science that allows us to judge relative properties of a class of objects like this,
which makes the periodic table a very useful tool. Many periodic trends are general. There may be a few points where an opposite
trend is seen, but there is an overall trend when considered across a whole row or down a whole column of the periodic table.
The first periodic trend we will consider is atomic radius. The atomic radius is an indication of the size of an atom. Although the
concept of a definite radius of an atom is a bit fuzzy, atoms behave as if they have a certain radius. Such radii can be estimated
from various experimental techniques, such as the x-ray crystallography of crystals.
As you go down a column of the periodic table, the atomic radii increase. This is because the valence electron shell is getting larger
and there is a larger principal quantum number, so the valence shell lies physically farther away from the nucleus. This trend can be
summarized as follows:
\[as\downarrow PT,atomic\; radius \uparrow \nonumber \]
where PT stands for periodic table. Going across a row on the periodic table, left to right, the trend is different. Even though the
valence shell maintains the same principal quantum number, the number of protons—and hence the nuclear charge—is increasing
as you go across the row. The increasing positive charge casts a tighter grip on the valence electrons, so as you go across the
periodic table, the atomic radii decrease. Again, we can summarize this trend as follows:

as → P T , atomic radius ↓

Figure 2.16.1 shows spheres representing the atoms of the s and p blocks from the periodic table to scale, showing the two trends
for the atomic radius.

Figure 2.16.1 : Atomic Radii Trends on the Periodic Table. Although there are some reversals in the trend (e.g., see Po in the
bottom row), atoms generally get smaller as you go across the periodic table and larger as you go down any one column. Numbers
are the radii in pm.

2.16.1 https://chem.libretexts.org/@go/page/423602
 Example 2.16.1: Atomic Radii

Referring only to a periodic table and not to Figure 2.16.1, which atom is larger in each pair?
a. Si or S
b. S or Te

Solution
a. Si is to the left of S on the periodic table; it is larger because as you go across the row, the atoms get smaller.
b. S is above Te on the periodic table; Te is larger because as you go down the column, the atoms get larger.

 Exercise 2.16.1: Atomic Radii

Referring only to a periodic table and not to Figure 2.16.1, which atom is smaller, Ca or Br?

Answer
Br

Ionization energy (IE) is the amount of energy required to remove an electron from an atom in the gas phase:
+ −
A(g) → A (g) + e ΔH ≡ I E

IE is usually expressed in kJ/mol of atoms. It is always positive because the removal of an electron always requires that energy be
put in (i.e., it is endothermic). IE also shows periodic trends. As you go down the periodic table, it becomes easier to remove an
electron from an atom (i.e., IE decreases) because the valence electron is farther away from the nucleus. Thus,

as ↓ P T , I E ↓

However, as you go across the periodic table and the electrons get drawn closer in, it takes more energy to remove an electron; as a
result, IE increases:

as → P T , I E ↑

Figure 2.16.2 shows values of IE versus position on the periodic table. Again, the trend is not absolute, but the general trends going
across and down the periodic table should be obvious.

2.16.2 https://chem.libretexts.org/@go/page/423602
Figure 2.16.2 : Ionization Energy on the Periodic Table. Values are in kJ/mol.
IE also shows an interesting trend within a given atom. This is because more than one IE can be defined by removing successive
electrons (if the atom has them to begin with):
First Ionization Energy (IE1):
+ −
A(g) → A (g) + e

Second Ionization Energy (IE2):


+ 2+ −
A (g) → A (g) + e

Third Ionization Energy (IE3):


2+ 3+ −
A (g) → A (g) + e

and so forth.
Each successive IE is larger than the previous because an electron is being removed from an atom with a progressively larger
positive charge. However, IE takes a large jump when a successive ionization goes down into a new shell. For example, the
following are the first three IEs for Mg, whose electron configuration is 1s22s22p63s2:
First Ionization Energy (IE1) = 738 kJ/mol:
+ −
M g(g) → M g (g) + e

Second Ionization Energy (IE2) = 1,450 kJ/mol:


+ 2+ −
Mg (g) → M g (g) + e

Third Ionization Energy (IE3) = 7,734 kJ/mol:


2+ 3+ −
Mg (g) → M g (g) + e

The second IE is twice the first, which is not a surprise: the first IE involves removing an electron from a neutral atom, while the
second one involves removing an electron from a positive ion. The third IE, however, is over five times the previous one. Why is it
so much larger? Because the first two electrons are removed from the 3s subshell, but the third electron has to be removed from the
n = 2 shell (specifically, the 2p subshell, which is lower in energy than the n = 3 shell). Thus, it takes much more energy than just
overcoming a larger ionic charge would suggest. It is trends like this that demonstrate that electrons within atoms are organized in
groups.

2.16.3 https://chem.libretexts.org/@go/page/423602
 Example 2.16.2: Ionization Energies
Which atom in each pair has the larger first ionization energy?
a. Ca or Sr
b. K or K+

Solution
a. Because Sr is below Ca on the periodic table, it is easier to remove an electron from it; thus, Ca has the higher IE.
b. Because K+ has a positive charge, it will be harder to remove another electron from it, so its IE is larger than that of K.
Indeed, it will be significantly larger because the next electron in K+ to be removed comes from another shell.

 Exercise 2.16.2: Ionization Energies

Which atom has the lower ionization energy, C or F?

Answer
C

The opposite of IE is described by electron affinity (EA), which is the energy change when a gas-phase atom accepts an electron:
\[A(g)+e^{-}\rightarrow A^{-}(g)\; \; \; \; \; \Delta H\equiv EA \nonumber \]
EA is also usually expressed in kJ/mol. EA also demonstrates some periodic trends, although they are less obvious than the other
periodic trends discussed previously. Generally, as you go across the periodic table, EA increases its magnitude:

as → P T , EA ↑

There is not a definitive trend as you go down the periodic table; sometimes EA increases, sometimes it decreases. Figure 2.16.3
shows EA values versus position on the periodic table for the s- and p-block elements. The trend is not absolute, especially
considering the large positive EA values for the second column. However, the general trend going across the periodic table should
be obvious.

2.16.4 https://chem.libretexts.org/@go/page/423602
Figure 2.16.3 : Electron Affinity on the Periodic Table. Values are in kJ/mol.

 Example 2.16.3: Electron Affinities

Predict which atom in each pair will have the highest magnitude of Electron Affinity.
a. C or F
b. Na or S

Solution
a. C and F are in the same row on the periodic table, but F is farther to the right. Therefore, F should have the larger
magnitude of EA.
b. Na and S are in the same row on the periodic table, but S is farther to the right. Therefore, S should have the larger
magnitude of EA.

 Exercise 2.16.3: Electron Affinities

Predict which atom will have the highest magnitude of Electron Affinity: As or Br.

Answer
Br

Metallic Character
The metallic character is used to define the chemical properties that metallic elements present. Generally, metals tend to lose
electrons to form cations. Nonmetals tend to gain electrons to form anions. They also have a high oxidation potential—therefore
they are easily oxidized and are strong reducing agents. Metals also form basic oxides; the more basic the oxide, the higher the
metallic character.

2.16.5 https://chem.libretexts.org/@go/page/423602
Figure 2.16.4 : Courtesy of Jessica Thornton (UCD)
As you move across the table from left to right, the metallic character decreases, because the elements easily accept electrons to fill
their valance shells. Therefore, these elements take on the nonmetallic character of forming anions. As you move up the table, the
metallic character decreases, due to the greater pull that the nucleus has on the outer electrons. This greater pull makes it harder for
the atoms to lose electrons and form cations.

 Uses of the Periodic Properties of Elements


1. Predict greater or smaller atomic size and radial distribution in neutral atoms and ions.
2. Measure and compare ionization energies.
3. Compare electron affinities and electronegativities.
Predict redox potential.
Compare metallic character with other elements; ability to form cations.
Predict reactions that may or may not occur due to the trends.
Determine greater cell potential (sum of oxidation and reduction potential) between reactions.
Complete chemical reactions according to trends.

Summary
Certain properties—notably atomic radius, ionization energies, and electron affinities—can be qualitatively understood by the
positions of the elements on the periodic table. The major trends are summarized in the figure below.
There are three factors that help in the prediction of the trends in the Periodic Table: number of protons in the nucleus, number
of shells, and shielding effect.

Various periodic trends (CC BY-SA 4.0; Sandbh via Wikipedia)

2.16: Periodic Trends - Atomic Size, Ionization Energy, and Metallic Character is shared under a CK-12 license and was authored, remixed,
and/or curated by Marisa Alviar-Agnew & Henry Agnew.
9.9: Periodic Trends - Atomic Size, Ionization Energy, and Metallic Character by Henry Agnew, Marisa Alviar-Agnew is licensed CK-
12. Original source: https://www.ck12.org/c/chemistry/.

2.16.6 https://chem.libretexts.org/@go/page/423602
2.17: Electron Configurations and the Periodic Table
 Learning Objectives
Relate the electron configurations of the elements to the shape of the periodic table.
Determine the expected electron configuration of an element by its place on the periodic table.

Previously, we introduced the periodic table as a tool for organizing the known chemical elements. A periodic table is shown in
Figure 2.17.1. The elements are listed by atomic number (the number of protons in the nucleus), and elements with similar
chemical properties are grouped together in columns.

Figure 2.17.1 : The Periodic Table


Why does the periodic table have the structure it does? The answer is rather simple, if you understand electron configurations: the
shape of the periodic table mimics the filling of the subshells with electrons.

The shape of the periodic table mimics the filling of the subshells with electrons.
Let us start with H and He. Their electron configurations are 1s1 and 1s2, respectively; with He, the n = 1 shell is filled. These two
elements make up the first row of the periodic table (Figure 2.17.2)

Figure 2.17.2 : The 1s Subshell. H and He represent the filling of the 1s subshell.
The next two electrons, for Li and Be, would go into the 2s subshell. Figure 2.17.3 shows that these two elements are adjacent on
the periodic table.

2.17.1 https://chem.libretexts.org/@go/page/423603
Figure 2.17.3 : The 2s Subshell. In Li and Be, the 2s subshell is being filled.
For the next six elements, the 2p subshell is being occupied with electrons. On the right side of the periodic table, these six
elements (B through Ne) are grouped together (Figure 2.17.4).

Figure 2.17.4 : The 2p Subshell. For B through Ne, the 2p subshell is being occupied.
The next subshell to be filled is the 3s subshell. The elements when this subshell is being filled, Na and Mg, are back on the left
side of the periodic table (Figure 2.17.5).

Figure 2.17.5 : The 3s Subshell. Now the 3s subshell is being occupied.


Next, the 3p subshell is filled with the next six elements (Figure 2.17.6).

2.17.2 https://chem.libretexts.org/@go/page/423603
Figure 2.17.6 : The 3p Subshell. Next, the 3p subshell is filled with electrons.
Instead of filling the 3d subshell next, electrons go into the 4s subshell (Figure 2.17.7).

Figure 2.17.7 : The 4s Subshell. The 4s subshell is filled before the 3d subshell. This is reflected in the structure of the periodic
table.
After the 4s subshell is filled, the 3d subshell is filled with up to 10 electrons. This explains the section of 10 elements in the
middle of the periodic table (Figure 2.17.8).

Figure 2.17.8 : The 3d Subshell. The 3d subshell is filled in the middle section of the periodic table.
...And so forth. As we go across the rows of the periodic table, the overall shape of the table outlines how the electrons are
occupying the shells and subshells.
The first two columns on the left side of the periodic table are where the s subshells are being occupied. Because of this, the first
two rows of the periodic table are labeled the s block. Similarly, the p block are the right-most six columns of the periodic table,
the d block is the middle 10 columns of the periodic table, while the f block is the 14-column section that is normally depicted as
detached from the main body of the periodic table. It could be part of the main body, but then the periodic table would be rather
long and cumbersome. Figure 2.17.9 shows the blocks of the periodic table.

2.17.3 https://chem.libretexts.org/@go/page/423603
Figure 2.17.9 : Blocks on the Periodic Table. The periodic table is separated into blocks depending on which subshell is being filled
for the atoms that belong in that section.
The electrons in the highest-numbered shell, plus any electrons in the last unfilled subshell, are called valence electrons; the
highest-numbered shell is called the valence shell. (The inner electrons are called core electrons.) The valence electrons largely
control the chemistry of an atom. If we look at just the valence shell’s electron configuration, we find that in each column, the
valence shell’s electron configuration is the same. For example, take the elements in the first column of the periodic table: H, Li,
Na, K, Rb, and Cs. Their electron configurations (abbreviated for the larger atoms) are as follows, with the valence shell electron
configuration highlighted:
Electrons, electron configurations, and the valence shell electron configuration highlighted.
H: 1s1

Li: 1s22s1

Na: [Ne]3s1

K: [Ar]4s1

Rb: [Kr]5s1

Cs: [Xe]6s1

They all have a similar electron configuration in their valence shells: a single s electron. Because much of the chemistry of an
element is influenced by valence electrons, we would expect that these elements would have similar chemistry—and they do. The
organization of electrons in atoms explains not only the shape of the periodic table, but also the fact that elements in the same
column of the periodic table have similar chemistry.
The same concept applies to the other columns of the periodic table. Elements in each column have the same valence shell electron
configurations, and the elements have some similar chemical properties. This is strictly true for all elements in the s and p blocks.
In the d and f blocks, because there are exceptions to the order of filling of subshells with electrons, similar valence shells are not
absolute in these blocks. However, many similarities do exist in these blocks, so a similarity in chemical properties is expected.
Similarity of valence shell electron configuration implies that we can determine the electron configuration of an atom solely by its
position on the periodic table. Consider Se, as shown in Figure 2.17.10. It is in the fourth column of the p block. This means that its
electron configuration should end in a p4 electron configuration. Indeed, the electron configuration of Se is [Ar]4s23d104p4, as
expected.

2.17.4 https://chem.libretexts.org/@go/page/423603
Figure 2.17.10: Selenium on the Periodic Table

 Example 2.17.1: Predicting Electron Configurations

From the element’s position on the periodic table, predict the valence shell electron configuration for each atom (Figure
2.17.11).

Figure 2.17.11: Various Elements on the Periodic Table


a. Ca
b. Sn

Solution
a. Ca is located in the second column of the s block. We expect that its electron configuration should end with s2. Calcium’s
electron configuration is [Ar]4s2.
b. Sn is located in the second column of the p block, so we expect that its electron configuration would end in p2. Tin’s
electron configuration is [Kr]5s24d105p2.

 Exercise 2.17.1

From the element’s position on the periodic table, predict the valence shell electron configuration for each atom. Figure
2.17.11.

a. Ti
b. Cl

Answer a
[Ar]4s23d2
Answer b
[Ne]3s23p5

2.17.5 https://chem.libretexts.org/@go/page/423603
Summary
The arrangement of electrons in atoms is responsible for the shape of the periodic table. Electron configurations can be
predicted by the position of an atom on the periodic table.

2.17: Electron Configurations and the Periodic Table is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa
Alviar-Agnew & Henry Agnew.
9.7: Electron Configurations and the Periodic Table by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

2.17.6 https://chem.libretexts.org/@go/page/423603
2.18: REPRESENTING VALENCE ELECTRONS WITH DOTS
 LEARNING OBJECTIVE
Draw a Lewis electron dot diagram for an atom or a monatomic ion.

In almost all cases, chemical bonds are formed by interactions of valence electrons in atoms. To facilitate our understanding of how valence
electrons interact, a simple way of representing those valence electrons would be useful.
A Lewis electron dot diagram (or electron dot diagram, or a Lewis diagram, or a Lewis structure) is a representation of the valence
electrons of an atom that uses dots around the symbol of the element. The number of dots equals the number of valence electrons in the
atom. These dots are arranged to the right and left and above and below the symbol, with no more than two dots on a side. (The order in
which the positions are used does not matter.) For example, the Lewis electron dot diagram for hydrogen is simply
H⋅

Because the side is not important, the Lewis electron dot diagram could also be drawn as follows:
˙
H or ⋅H or H
.

The electron dot diagram for helium, with two valence electrons, is as follows:
He:

By putting the two electrons together on the same side, we emphasize the fact that these two electrons are both in the 1s subshell; this is the
common convention we will adopt, although there will be exceptions later. The next atom, lithium, has an electron configuration of 1s22s1,
so it has only one electron in its valence shell. Its electron dot diagram resembles that of hydrogen, except the symbol for lithium is used:

Li⋅

Beryllium has two valence electrons in its 2s shell, so its electron dot diagram is like that of helium:

Be:

The next atom is boron. Its valence electron shell is 2s22p1, so it has three valence electrons. The third electron will go on another side of
the symbol:

Ḃ:

Again, it does not matter on which sides of the symbol the electron dots are positioned.
For carbon, there are four valence electrons, two in the 2s subshell and two in the 2p subshell. As usual, we will draw two dots together on
one side, to represent the 2s electrons. However, conventionally, we draw the dots for the two p electrons on different sides. As such, the
electron dot diagram for carbon is as follows:
˙
⋅C:

With N, which has three p electrons, we put a single dot on each of the three remaining sides:
˙
⋅N:
.

For oxygen, which has four p electrons, we now have to start doubling up on the dots on one other side of the symbol. When doubling up
electrons, make sure that each side has no more than two electrons.
¨
⋅O:
.

Fluorine and neon have seven and eight dots, respectively:


¨:
:F
.

¨
:Ne:
. .

With the next element, sodium, the process starts over with a single electron because sodium has a single electron in its highest-numbered
shell, the n = 3 shell. By going through the periodic table, we see that the Lewis electron dot diagrams of atoms will never have more than

2.18.1 https://chem.libretexts.org/@go/page/423604
eight dots around the atomic symbol.

 EXAMPLE 2.18.1: LEWIS DOT DIAGRAMS

What is the Lewis electron dot diagram for each element?


a. aluminum
b. selenium

Solution
a. The valence electron configuration for aluminum is 3s23p1. So it would have three dots around the symbol for aluminum, two of
them paired to represent the 3s electrons:
˙
Al :

2. The valence electron configuration for selenium is 4s24p4. In the highest-numbered shell, the n = 4 shell, there are six electrons. Its
electron dot diagram is as follows:
˙
⋅S e:
. .

 EXERCISE 2.18.1

What is the Lewis electron dot diagram for each element?


a. phosphorus
b. argon

Answer a
˙
⋅P:
.

Answer b
¨
:Ar:
. .

SUMMARY
Lewis electron dot diagrams use dots to represent valence electrons around an atomic symbol.
Lewis electron dot diagrams for ions have less (for cations) or more (for anions) dots than the corresponding atom.

2.18: Representing Valence Electrons with Dots is shared under a CC BY-NC-SA 3.0 license and was authored, remixed, and/or curated by Marisa Alviar-
Agnew & Henry Agnew.
10.2: Representing Valence Electrons with Dots by Henry Agnew, Marisa Alviar-Agnew is licensed CC BY-NC-SA 3.0.

2.18.2 https://chem.libretexts.org/@go/page/423604
2.19: LEWIS STRUCTURES OF IONIC COMPOUNDS- ELECTRONS
TRANSFERRED
 LEARNING OBJECTIVES
State the octet rule.
Define ionic bond.
Draw Lewis structures for ionic compounds.

In Section 4.7, we demonstrated that ions are formed by losing electrons to make cations, or by gaining electrons to form anions. The astute
reader may have noticed something: many of the ions that form have eight electrons in their valence shell. Either atoms gain enough
electrons to have eight electrons in the valence shell and become the appropriately charged anion, or they lose the electrons in their original
valence shell; the lower shell, now the valence shell, has eight electrons in it, so the atom becomes positively charged. For whatever reason,
having eight electrons in a valence shell is a particularly energetically stable arrangement of electrons. The octet rule explains the favorable
trend of atoms having eight electrons in their valence shell. When atoms form compounds, the octet rule is not always satisfied for all atoms
at all times, but it is a very good rule of thumb for understanding the kinds of bonding arrangements that atoms can make.
It is not impossible to violate the octet rule. Consider sodium: in its elemental form, it has one valence electron and is stable. It is rather
reactive, however, and does not require a lot of energy to remove that electron to make the Na+ ion. We could remove another electron by
adding even more energy to the ion, to make the Na2+ ion. However, that requires much more energy than is normally available in chemical
reactions, so sodium stops at a 1+ charge after losing a single electron. It turns out that the Na+ ion has a complete octet in its new valence
shell, the n = 2 shell, which satisfies the octet rule. The octet rule is a result of trends in energies and is useful in explaining why atoms form
the ions that they do.
Now consider an Na atom in the presence of a Cl atom. The two atoms have these Lewis electron dot diagrams and electron configurations:
¨
Na ⋅ ⋅Cl :
. .

1 2 5
[N e] 3s [N e] 3s 3p

For the Na atom to obtain an octet, it must lose an electron; for the Cl atom to gain an octet, it must gain an electron. An electron transfers
from the Na atom to the Cl atom:
¨l :
Na ⋅ ↷ ⋅C
. .

resulting in two ions—the Na+ ion and the Cl− ion:


+ ¨ −
Na :Cl :
. .

2 6
[N e] [N e] 3s 3p

Both species now have complete octets, and the electron shells are energetically stable. From basic physics, we know that opposite charges
attract. This is what happens to the Na+ and Cl− ions:
+ ¨ − + −
Na + :Cl : → N a Cl or N aCl
. .

where we have written the final formula (the formula for sodium chloride) as per the convention for ionic compounds, without listing the
charges explicitly. The attraction between oppositely charged ions is called an ionic bond, and it is one of the main types of chemical bonds
in chemistry. Ionic bonds are caused by electrons transferring from one atom to another.
In electron transfer, the number of electrons lost must equal the number of electrons gained. We saw this in the formation of NaCl. A similar
process occurs between Mg atoms and O atoms, except in this case two electrons are transferred:

The two ions each have octets as their valence shell, and the two oppositely charged particles attract, making an ionic bond:
2−
2+ ¨ 2+ 2−
Mg + [:O :] Mg O or M gO
. .

2.19.1 https://chem.libretexts.org/@go/page/423605
Remember, in the final formula for the ionic compound, we do not write the charges on the ions.
What about when an Na atom interacts with an O atom? The O atom needs two electrons to complete its valence octet, but the Na atom
supplies only one electron:
¨
Na ⋅ ↷ ⋅O :
.

The O atom still does not have an octet of electrons. What we need is a second Na atom to donate a second electron to the O atom:

These three ions attract each other to give an overall neutral-charged ionic compound, which we write as Na2O. The need for the number of
electrons lost being equal to the number of electrons gained explains why ionic compounds have the ratio of cations to anions that they do.
This is required by the law of conservation of matter as well.

 EXAMPLE 2.19.1: SYNTHESIS OF CALCIUM CHLORIDE FROM ELEMENTS

With arrows, illustrate the transfer of electrons to form calcium chloride from Ca atoms and Cl atoms.

Solution
A Ca atom has two valence electrons, while a Cl atom has seven electrons. A Cl atom needs only one more to complete its octet,
while Ca atoms have two electrons to lose. Thus we need two Cl atoms to accept the two electrons from one Ca atom. The transfer
process looks as follows:

The oppositely charged ions attract each other to make CaCl2.

 EXERCISE 2.19.1

With arrows, illustrate the transfer of electrons to form potassium sulfide from K atoms and S atoms.

Answer

SUMMARY
The tendency to form species that have eight electrons in the valence shell is called the octet rule.
The attraction of oppositely charged ions caused by electron transfer is called an ionic bond.
The strength of ionic bonding depends on the magnitude of the charges and the sizes of the ions.

2.19: Lewis Structures of Ionic Compounds- Electrons Transferred is shared under a CC BY-NC-SA 3.0 license and was authored, remixed, and/or curated
by Marisa Alviar-Agnew & Henry Agnew.
10.3: Lewis Structures of Ionic Compounds- Electrons Transferred by Henry Agnew, Marisa Alviar-Agnew is licensed CC BY-NC-SA 3.0.

2.19.2 https://chem.libretexts.org/@go/page/423605
2.20: Covalent Lewis Structures- Electrons Shared
 Learning Objectives
Define covalent bond.
Illustrate covalent bond formation with Lewis electron dot diagrams.

Ionic bonding typically occurs when it is easy for one atom to lose one or more electrons and another atom to gain one or more
electrons. However, some atoms won’t give up or gain electrons easily. Yet they still participate in compound formation. How?
There is another mechanism for obtaining a complete valence shell: sharing electrons. When electrons are shared between two
atoms, they make a bond called a covalent bond.
Let us illustrate a covalent bond by using H atoms, with the understanding that H atoms need only two electrons to fill the 1s
subshell. Each H atom starts with a single electron in its valence shell:

H⋅ ⋅ H

The two H atoms can share their electrons:

H : H

We can use circles to show that each H atom has two electrons around the nucleus, completely filling each atom’s valence shell:

Because each H atom has a filled valence shell, this bond is stable, and we have made a diatomic hydrogen molecule. (This
explains why hydrogen is one of the diatomic elements.) For simplicity’s sake, it is not unusual to represent the covalent bond with
a dash, instead of with two dots:
H–H
Because two atoms are sharing one pair of electrons, this covalent bond is called a single bond. As another example, consider
fluorine. F atoms have seven electrons in their valence shell:

These two atoms can do the same thing that the H atoms did; they share their unpaired electrons to make a covalent bond.

Note that each F atom has a complete octet around it now:

We can also write this using a dash to represent the shared electron pair:

There are two different types of electrons in the fluorine diatomic molecule. The bonding electron pair makes the covalent bond.
Each F atom has three other pairs of electrons that do not participate in the bonding; they are called lone pair electrons. Each F
atom has one bonding pair and three lone pairs of electrons.

2.20.1 https://chem.libretexts.org/@go/page/423606
Covalent bonds can be made between different elements as well. One example is HF. Each atom starts out with an odd number of
electrons in its valence shell:

The two atoms can share their unpaired electrons to make a covalent bond:

We note that the H atom has a full valence shell with two electrons, while the F atom has a complete octet of electrons.

 Example 2.20.1:

Use Lewis electron dot diagrams to illustrate the covalent bond formation in HBr.

Solution
HBr is very similar to HF, except that it has Br instead of F. The atoms are as follows:

The two atoms can share their unpaired electron:

 Exercise 2.20.1

Use Lewis electron dot diagrams to illustrate the covalent bond formation in Cl2.

Answer

When working with covalent structures, it sometimes looks like you have leftover electrons. You apply the rules you learned so far,
and there are still some electrons that remain unattached. You can't just leave them there. So where do you put them?

Multiple Covalent Bonds


Some molecules are not able to satisfy the octet rule by making only single covalent bonds between the atoms. Consider the
compound ethene, which has a molecular formula of C H . The carbon atoms are bonded together, with each carbon also bonded
2 4

to two hydrogen atoms.


two C atoms = 2 × 4 = 8 valence electrons
four H atoms = 4 × 1 = 4 valence electrons
total of 12 valence electrons in the molecule
If the Lewis electron dot structure was drawn with a single bond between the carbon atoms and with the octet rule followed, it
would look like this:

2.20.2 https://chem.libretexts.org/@go/page/423606
Figure 2.20.1 : Incorrect dot structure of ethene. (CK12 License)
This Lewis structure is incorrect because it contains a total of 14 electrons. However, the Lewis structure can be changed by
eliminating the lone pairs on the carbon atoms and having to share two pairs instead of only one pair.

Figure 2.20.2 : Correct dot structure for ethene. (CK12 License)


A double covalent bond is a covalent bond formed by atoms that share two pairs of electrons. The double covalent bond that
occurs between the two carbon atoms in ethane can also be represented by a structural formula and with a molecular model as
shown in the figure below.

Figure 2.20.3 : (A) The structural model for C H consists of a double covalent bond between the two carbon atoms and single
2 4

bonds to the hydrogen atoms. (B) Molecular model of C H .


2 4

A triple covalent bond is a covalent bond formed by atoms that share three pairs of electrons. The element nitrogen is a gas that
composes the majority of Earth's atmosphere. A nitrogen atom has five valence electrons, which can be shown as one pair and three
single electrons. When combining with another nitrogen atom to form a diatomic molecule, the three single electrons on each atom
combine to form three shared pairs of electrons.

Figure 2.20.4 : Triple bond in N .


2

Each nitrogen atom follows the octet rule with one lone pair of electrons, and six electrons that are shared between the atoms.

Summary
Covalent bonds are formed when atoms share electrons.
Lewis electron dot diagrams can be drawn to illustrate covalent bond formation.
Double bonds or triple bonds between atoms may be necessary to properly illustrate the bonding in some molecules.

Contributions & Attributions

Anonymous by request

2.20: Covalent Lewis Structures- Electrons Shared is shared under a mixed license and was authored, remixed, and/or curated by Marisa Alviar-
Agnew & Henry Agnew.
10.4: Covalent Lewis Structures- Electrons Shared by Henry Agnew, Marisa Alviar-Agnew is licensed notset.

2.20.3 https://chem.libretexts.org/@go/page/423606
2.21: Writing Lewis Structures for Covalent Compounds
 Learning Objectives
Draw Lewis structures for covalent compounds.

The following procedure can be used to construct Lewis electron structures for more complex molecules and ions.

 How-to: Constructing Lewis electron structures

1. Determine the total number of valence electrons in the molecule or ion.


Add together the valence electrons from each atom. (Recall that the number of valence electrons is indicated by the position
of the element in the periodic table.)
If the species is a polyatomic ion, remember to add or subtract the number of electrons necessary to give the total charge on
the ion.
For CO32−, for example, we add two electrons to the total because of the −2 charge.

2. Arrange the atoms to show specific connections.


When there is a central atom, it is usually the least electronegative element in the compound. Chemists usually list this
central atom first in the chemical formula (as in CCl4 and CO32−, which both have C as the central atom), which is another
clue to the compound’s structure.
Hydrogen and the halogens are almost always connected to only one other atom, so they are usually terminal rather than
central.

3. Place a bonding pair of electrons between each pair of adjacent atoms to give a single bond.
In H2O, for example, there is a bonding pair of electrons between oxygen and each hydrogen.
4. Beginning with the terminal atoms, add enough electrons to each atom to give each atom an octet (two for hydrogen).
These electrons will usually be lone pairs.
5. If any electrons are left over, place them on the central atom.
We will explain later that some atoms are able to accommodate more than eight electrons.

6. If the central atom has fewer electrons than an octet, use lone pairs from terminal atoms to
form multiple (double or triple) bonds to the central atom to achieve an octet.
This will not change the number of electrons on the terminal atoms.

7. Final check
Always make sure all valence electrons are accounted for and that each atom has an octet of electrons, except for hydrogen
(with two electrons).
The central atom is usually the least electronegative element in the molecule or ion; hydrogen and the halogens are usually
terminal.

Now let’s apply this procedure to some particular compounds, beginning with one we have already discussed.

2.21.1 https://chem.libretexts.org/@go/page/423607
 Example 2.21.1: Water

Write the Lewis Structure for H2O.

Solution
Solutions to Example 10.4.1
Steps for Writing Lewis Structures Example 2.21.1

1. Determine the total number of valence electrons in the molecule or Each H atom (group 1) has 1 valence electron, and the O atom (group
ion. 16) has 6 valence electrons, for a total of 8 valence electrons.

2. Arrange the atoms to show specific connections.


HOH
Because H atoms are almost always terminal, the arrangement within
the molecule must be HOH.
Placing one bonding pair of electrons between the O atom and each H
3. Place a bonding pair of electrons between each pair of adjacent atom gives
atoms to give a single bond.
4. Beginning with the terminal atoms, add enough electrons to each H -O- H
atom to give each atom an octet (two for hydrogen). with 4 electrons left over.
Each H atom has a full valence shell of 2 electrons.
Adding the remaining 4 electrons to the oxygen (as two lone pairs)
gives the following structure:
5. If any electrons are left over, place them on the central atom.

6. If the central atom has fewer electrons than an octet, use lone pairs
from terminal atoms to form multiple (double or triple) bonds to the Not necessary.
central atom to achieve an octet.

The Lewis structure gives oxygen an octet and each hydrogen 2


7. Final check.
electrons.

 Example 2.21.2

Write the Lewis structure for the C H 2O molecule

Solution
Solutions to Example 10.4.2
Steps for Writing Lewis Structures Example 2.21.2

Each hydrogen atom (group 1) has 1 valence electron, carbon (group


1. Determine the total number of valence electrons in the molecule or
14) has 4 valence electrons, and oxygen (group 16) has 6 valence
ion.
electrons, for a total of [(2)(1) + 4 + 6] = 12 valence electrons.

2. Arrange the atoms to show specific connections.

Because carbon is less electronegative than oxygen and hydrogen is


normally terminal, C must be the central atom.

2.21.2 https://chem.libretexts.org/@go/page/423607
Steps for Writing Lewis Structures Example 2.21.2

Placing a bonding pair of electrons between each pair of bonded


atoms gives the following:

3. Place a bonding pair of electrons between each pair of adjacent


atoms to give a single bond.

6 electrons are used, and 6 are left over.


Adding all 6 remaining electrons to oxygen (as three lone pairs) gives
the following:

4. Beginning with the terminal atoms, add enough electrons to each


atom to give each atom an octet (two for hydrogen).

Although oxygen now has an octet and each hydrogen has 2


electrons, carbon has only 6 electrons.
Not necessary.
5. If any electrons are left over, place them on the central atom.
There are no electrons left to place on the central atom.
To give carbon an octet of electrons, we use one of the lone pairs of
6. If the central atom has fewer electrons than an octet, use lone pairs electrons on oxygen to form a carbon–oxygen double bond:
from terminal atoms to form multiple (double or triple) bonds to the
central atom to achieve an octet.

Both the oxygen and the carbon now have an octet of electrons, so
this is an acceptable Lewis electron structure. The O has two bonding
7. Final check
pairs and two lone pairs, and C has four bonding pairs. This is the
structure of formaldehyde, which is used in embalming fluid.

 Exercise 2.21.1

Write Lewis electron structures for CO2 and SCl2, a vile-smelling, unstable red liquid that is used in the manufacture of rubber.

Answer CO2

Answer SCl2

The United States Supreme Court has the unenviable task of deciding what the law is. This responsibility can be a major challenge
when there is no clear principle involved or where there is a new situation not encountered before. Chemistry faces the same
challenge in extending basic concepts to fit a new situation. Drawing of Lewis structures for polyatomic ions uses the same
approach, but tweaks the process a little to fit a somewhat different set of circumstances.

2.21.3 https://chem.libretexts.org/@go/page/423607
Writing Lewis Structures for Polyatomic Ions (CK-12)
Recall that a polyatomic ion is a group of atoms that are covalently bonded together and which carry an overall electrical charge.
The ammonium ion, NH , is formed when a hydrogen ion (H ) attaches to the lone pair of an ammonia (NH ) molecule in a
+

4
+

coordinate covalent bond.

Figure 2.21.3 : The ammonium ion. (CK12 License)


When drawing the Lewis structure of a polyatomic ion, the charge of the ion is reflected in the number of total valence electrons in
the structure. In the case of the ammonium ion:
1 N atom = 5 valence electrons
4 H atoms = 4 × 1 = 4 valence electrons
subtract 1 electron for the 1+ charge of the ion
total of 8 valence electrons in the ion
It is customary to put the Lewis structure of a polyatomic ion into a large set of brackets, with the charge of the ion as a superscript
outside of the brackets.

 Exercise 2.21.2

Draw the Lewis electron dot structure for the sulfate ion.

Answer (CK12 License)

Exceptions to the Octet Rule (BC Campus)


As important and useful as the octet rule is in chemical bonding, there are some well-known violations. This does not mean that the
octet rule is useless—quite the contrary. As with many rules, there are exceptions, or violations.
There are three violations to the octet rule. Odd-electron molecules represent the first violation to the octet rule. Although they are
few, some stable compounds have an odd number of electrons in their valence shells. With an odd number of electrons, at least one
atom in the molecule will have to violate the octet rule. Examples of stable odd-electron molecules are NO, NO2, and ClO2. The
Lewis electron dot diagram for NO is as follows:

Although the O atom has an octet of electrons, the N atom has only seven electrons in its valence shell. Although NO is a stable
compound, it is very chemically reactive, as are most other odd-electron compounds.
Electron-deficient molecules represent the second violation to the octet rule. These stable compounds have less than eight electrons
around an atom in the molecule. The most common examples are the covalent compounds of beryllium and boron. For example,
beryllium can form two covalent bonds, resulting in only four electrons in its valence shell:

2.21.4 https://chem.libretexts.org/@go/page/423607
Boron commonly makes only three covalent bonds, resulting in only six valence electrons around the B atom. A well-known
example is BF3:

The third violation to the octet rule is found in those compounds with more than eight electrons assigned to their valence shell.
These are called expanded valence shell molecules. Such compounds are formed only by central atoms in the third row of the
periodic table or beyond that have empty d orbitals in their valence shells that can participate in covalent bonding. One such
compound is PF5. The only reasonable Lewis electron dot diagram for this compound has the P atom making five covalent bonds:

Formally, the P atom has 10 electrons in its valence shell.

 Example 2.21.3: Octet Violations

Identify each violation to the octet rule by drawing a Lewis electron dot diagram.
a. ClO
b. SF6

Solution
a. With one Cl atom and one O atom, this molecule has 6 + 7 = 13 valence electrons, so it is an odd-electron molecule. A
Lewis electron dot diagram for this molecule is as follows:
alt

b. In SF6, the central S atom makes six covalent bonds to the six surrounding F atoms, so it is an expanded valence shell
molecule. Its Lewis electron dot diagram is as follows:
alt

 Exercise 2.21.3: Xenon Difluoride

Identify the violation to the octet rule in XeF2 by drawing a Lewis electron dot diagram.

Answer

The Xe atom has an expanded valence shell with more than eight electrons around it.

Summary
Lewis dot symbols provide a simple rationalization of why elements form compounds with the observed stoichiometries. A plot of
the overall energy of a covalent bond as a function of internuclear distance is identical to a plot of an ionic pair because both result

2.21.5 https://chem.libretexts.org/@go/page/423607
from attractive and repulsive forces between charged entities. In Lewis electron structures, we encounter bonding pairs, which are
shared by two atoms, and lone pairs, which are not shared between atoms. Lewis structures for polyatomic ions follow the same
rules as those for other covalent compounds. There are three violations to the octet rule: odd-electron molecules, electron-deficient
molecules, and expanded valence shell molecules.

2.21: Writing Lewis Structures for Covalent Compounds is shared under a mixed license and was authored, remixed, and/or curated by Marisa
Alviar-Agnew & Henry Agnew.
10.5: Writing Lewis Structures for Covalent Compounds by Henry Agnew, Marisa Alviar-Agnew is licensed notset. Original source:
https://www.ck12.org/c/chemistry/.

2.21.6 https://chem.libretexts.org/@go/page/423607
2.21.7 https://chem.libretexts.org/@go/page/423607
2.21.8 https://chem.libretexts.org/@go/page/423607
2.22: PREDICTING THE SHAPES OF MOLECULES
 LEARNING OBJECTIVE
Determine the shape of simple molecules.

Molecules have shapes. There is an abundance of experimental evidence to that effect—from their physical properties to their chemical
reactivity. Small molecules—molecules with a single central atom—have shapes that can be easily predicted. The basic idea in molecular
shapes is called valence shell electron pair repulsion (VSEPR). VSEPR says that electron pairs, being composed of negatively charged
particles, repel each other to get as far away from one another as possible. VSEPR makes a distinction between electron group geometry,
which expresses how electron groups (bonds and nonbonding electron pairs) are arranged, and molecular geometry, which expresses how
the atoms in a molecule are arranged. However, the two geometries are related.
There are two types of electron groups: any type of bond—single, double, or triple—and lone electron pairs. When applying VSEPR to
simple molecules, the first thing to do is to count the number of electron groups around the central atom. Remember that a multiple bond
counts as only one electron group.
Any molecule with only two atoms is linear. A molecule whose central atom contains only two electron groups orients those two groups as
far apart from each other as possible—180° apart. When the two electron groups are 180° apart, the atoms attached to those electron groups
are also 180° apart, so the overall molecular shape is linear. Examples include BeH2 and CO2:

Figure 2.22.1: Beryllium hydride and carbon dioxide bonding.


The two molecules, shown in the figure below in a "ball and stick" model.

Figure 2.22.2: Beryllium hydride and carbon dioxide models. (CK12 Licence)
A molecule with three electron groups orients the three groups as far apart as possible. They adopt the positions of an equilateral triangle—
120° apart and in a plane. The shape of such molecules is trigonal planar. An example is BF3:

Figure 2.22.3: Boron trifluoride bonding. (CK12 Licence)


Some substances have a trigonal planar electron group distribution but have atoms bonded to only two of the three electron groups. An
example is GeF2:

Figure 2.22.4: Germanium difluoride bonding.


From an electron group geometry perspective, GeF2 has a trigonal planar shape, but its real shape is dictated by the positions of the atoms.
This shape is called bent or angular.
A molecule with four electron groups about the central atom orients the four groups in the direction of a tetrahedron, as shown in Figure
2.22.1 Tetrahedral Geometry. If there are four atoms attached to these electron groups, then the molecular shape is also tetrahedral.

Methane (CH4) is an example.

2.22.1 https://chem.libretexts.org/@go/page/423608
Figure 2.22.5: Tetrahedral structure of methane. (CK12 Licence)
This diagram of CH4 illustrates the standard convention of displaying a three-dimensional molecule on a two-dimensional surface. The
straight lines are in the plane of the page, the solid wedged line is coming out of the plane toward the reader, and the dashed wedged line is
going out of the plane away from the reader.

Figure 2.22.6: Methane bonding. (CK12 Licence)


NH3 is an example of a molecule whose central atom has four electron groups, but only three of them are bonded to surrounding atoms.

Figure 2.22.7: Ammonia bonding. (CK12 Licence)


Although the electron groups are oriented in the shape of a tetrahedron, from a molecular geometry perspective, the shape of NH3 is
trigonal pyramidal.
H2O is an example of a molecule whose central atom has four electron groups, but only two of them are bonded to surrounding atoms.

Figure 2.22.8: Water bonding.


Although the electron groups are oriented in the shape of a tetrahedron, the shape of the molecule is bent or angular. A molecule with four
electron groups about the central atom, but only one electron group bonded to another atom, is linear because there are only two atoms in
the molecule.
Double or triple bonds count as a single electron group. The Lewis electron dot diagram of formaldehyde (CH2O) is shown in Figure
2.22.9 .

Figure 2.22.9: Lewis Electron Dot Diagram of Formaldehyde.


The central C atom has three electron groups around it because the double bond counts as one electron group. The three electron groups
repel each other to adopt a trigonal planar shape.

Figure 2.22.10: Formaldehyde bonding.


(The lone electron pairs on the O atom are omitted for clarity.) The molecule will not be a perfect equilateral triangle because the C–O
double bond is different from the two C–H bonds, but both planar and triangular describe the appropriate approximate shape of this
molecule.
Table 2.22.1 summarizes the shapes of molecules based on the number of electron groups and surrounding atoms.

2.22.2 https://chem.libretexts.org/@go/page/423608
Table 2.22.1: Summary of Molecular Shapes
Number of Electron Groups on Central Atom Number of Bonding Groups Number of Lone Pairs Electron Geometry Molecular Shape
2 2 0 linear linear
3 3 0 trigonal planar trigonal planar
3 2 1 trigonal planar bent
4 4 0 tetrahedral tetrahedral
4 3 1 tetrahedral trigonal pyramidal
4 2 2 tetrahedral bent

 EXAMPLE 2.22.1

What is the approximate shape of each molecule?


a. PCl3
b. NOF

Solution
The first step is to draw the Lewis structure of the molecule.
For PCl , the electron dot diagram is as follows:
3

The lone electron pairs on the Cl atoms are omitted for clarity. The P atom has four electron groups with three of them bonded to
surrounding atoms, so the molecular shape is trigonal pyramidal.
The electron dot diagram for NOF is as follows:

The N atom has three electron groups on it, two of which are bonded to other atoms. The molecular shape is bent.

 EXERCISE 2.22.1

What is the approximate molecular shape of CH 2


Cl
2
?

Answer
Tetrahedral

 EXERCISE 2.22.2

Ethylene (C H ) has two central atoms. Determine the geometry around each central atom and the shape of the overall molecule.
2 4

(Hint: hydrogen is a terminal atom.)

Answer
Trigonal planar about both central C atoms.

SUMMARY
The approximate shape of a molecule can be predicted from the number of electron groups and the number of surrounding atoms.

2.22: Predicting the Shapes of Molecules is shared under a mixed 3.0 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew & Henry
Agnew.
10.7: Predicting the Shapes of Molecules by Henry Agnew, Marisa Alviar-Agnew is licensed notset.

2.22.3 https://chem.libretexts.org/@go/page/423608
2.23: Electronegativity and Polarity - Why Oil and Water Do not Mix
 Learning Objectives
Explain how polar compounds differ from nonpolar compounds.
Determine if a molecule is polar or nonpolar.
Given a pair of compounds, predict which would have a higher melting or boiling point.

Bond Polarity
The ability of an atom in a molecule to attract shared electrons is called electronegativity. When two atoms combine, the
difference between their electronegativities is an indication of the type of bond that will form. If the difference between the
electronegativities of the two atoms is small, neither atom can take the shared electrons completely away from the other atom, and
the bond will be covalent. If the difference between the electronegativities is large, the more electronegative atom will take the
bonding electrons completely away from the other atom (electron transfer will occur), and the bond will be ionic. This is why
metals (low electronegativities) bonded with nonmetals (high electronegativities) typically produce ionic compounds.
A bond may be so polar that an electron actually transfers from one atom to another, forming a true ionic bond. How do we judge
the degree of polarity? Scientists have devised a scale called electronegativity, a scale for judging how much atoms of any element
attract electrons. Electronegativity is a unitless number; the higher the number, the more an atom attracts electrons. A common
scale for electronegativity is shown in Figure 2.23.1.

Figure 2.23.1 : Electronegativities of the Elements. Electronegativities are used to determine the polarity of covalent bonds.
The polarity of a covalent bond can be judged by determining the difference of the electronegativities of the two atoms involved in
the covalent bond, as summarized in the following table:
difference of the electronegativities of the two atoms involved in the covalent bond
Electronegativity Difference Bond Type

0–0.4 pure covalent

0.5–2.0 polar covalent

>2.0 likely ionic

Nonpolar Covalent Bonds


A bond in which the electronegativity difference is less than 1.9 is considered to be mostly covalent in character. However, at this
point we need to distinguish between two general types of covalent bonds. A nonpolar covalent bond is a covalent bond in which
the bonding electrons are shared equally between the two atoms. In a nonpolar covalent bond, the distribution of electrical charge is
balanced between the two atoms.

2.23.1 https://chem.libretexts.org/@go/page/423609
Figure 2.23.2 : A nonpolar covalent bond is one in which the distribution of electron density between the two atoms is equal.
The two chlorine atoms share the pair of electrons in the single covalent bond equally, and the electron density surrounding the Cl 2

molecule is symmetrical. Also note that molecules in which the electronegativity difference is very small (<0.5) are also considered
nonpolar covalent. An example would be a bond between chlorine and bromine (ΔEN = 3.0 − 2.8 = 0.2 ).

Polar Covalent Bonds


A bond in which the electronegativity difference between the atoms is between 0.5 and 2.0 is called a polar covalent bond. A polar
covalent bond is a covalent bond in which the atoms have an unequal attraction for electrons and so the sharing is unequal. In a
polar covalent bond, sometimes simply called a polar bond, the distribution of electrons around the molecule is no longer
symmetrical.

Figure 2.23.3 : In the polar covalent bond of HF, the electron density is unevenly distributed. There is a higher density (red) near
the fluorine atom, and a lower density (blue) near the hydrogen atom.
An easy way to illustrate the uneven electron distribution in a polar covalent bond is to use the Greek letter delta (δ).

Figure 2.23.4 : Use of δ to indicate partial charge.


The atom with the greater electronegativity acquires a partial negative charge, while the atom with the lesser electronegativity
acquires a partial positive charge. The delta symbol is used to indicate that the quantity of charge is less than one. A crossed arrow
can also be used to indicate the direction of greater electron density.

Figure 2.23.5 : Use of crossed arrow to indicate polarity.

2.23.2 https://chem.libretexts.org/@go/page/423609
Electronegativity and bonding | Chemic…
Chemic…

Electronegativity differences in bonding using Pauling scale. Differences in electronegativity classify bonds as covalent, polar
covalent, or ionic.

 Example 2.23.1: Bond Polarity

What is the polarity of each bond?


a. C–H
b. O–H

Solution
Using Figure 2.23.1, we can calculate the difference of the electronegativities of the atoms involved in the bond.
a. For the C–H bond, the difference in the electronegativities is 2.5 − 2.1 = 0.4. Thus we predict that this bond will be
nonpolar covalent.
b. For the O–H bond, the difference in electronegativities is 3.5 − 2.1 = 1.4, so we predict that this bond will be polar
covalent.

 Exercise 2.23.1

What is the polarity of each bond?


a. Rb–F
b. P–Cl

Answer a
likely ionic
Answer b
polar covalent

Molecular Polarity
To determine if a molecule is polar or nonpolar, it is generally useful to look at Lewis structures. Nonpolar compounds will be
symmetric, meaning all of the sides around the central atom are identical—bonded to the same element with no unshared pairs of
electrons. Polar molecules are asymmetric, either containing lone pairs of electrons on a central atom or having atoms with
different electronegativities bonded. This works pretty well, as long as you can visualize the molecular geometry. That's the hard
part. To know how the bonds are oriented in space, you have to have a strong grasp of Lewis structures and VSEPR theory.
Assuming that you do, you can look at the structure of each one and decide if it is polar or not, whether or not you know the

2.23.3 https://chem.libretexts.org/@go/page/423609
individual atom's electronegativity. This is because you know that all bonds between dissimilar elements are polar, and in these
particular examples, it doesn't matter which direction the dipole moment vectors are pointing (out or in).
A polar molecule is a molecule in which one end of the molecule is slightly positive, while the other end is slightly negative. A
diatomic molecule that consists of a polar covalent bond, such as HF, is a polar molecule. The two electrically charged regions on
either end of the molecule are called poles, similar to a magnet having a north and a south pole. A molecule with two poles is called
a dipole (see figure below). Hydrogen fluoride is a dipole.

Figure 2.23.6 : A dipole is any molecule with a positive end and a negative end, resulting from unequal distribution of electron
density throughout the molecule.
For molecules with more than two atoms, the molecular geometry must also be taken into account when determining if the
molecule is polar or nonpolar. The figure below shows a comparison between carbon dioxide and water. Carbon dioxide (CO ) is a 2

linear molecule. The oxygen atoms are more electronegative than the carbon atom, so there are two individual dipoles pointing
outward from the C atom to each O atom. However, since the dipoles are of equal strength and are oriented this way, they cancel
out and the overall molecular polarity of CO is zero.
2

Water is a bent molecule because of the two lone pairs on the central oxygen atom. The individual dipoles point from the H atoms
toward the O atom. Because of the shape, the dipoles do not cancel each other out and the water molecule is polar. In the figure
below, the net dipole is shown in blue and points upward.

Figure 2.23.7 : The molecular geometry of a molecule affects its polarity. Each CO bond has a dipole moment, but they point in
opposite directions so that the net CO2 molecule is nonpolar. In contrast, water is polar because the OH bond moments do not
cancel out.
Some other molecules are shown in the figure below. Notice that a tetrahedral molecule such as CH is nonpolar. However, if one
4

of the peripheral H atoms is replaced with another atom that has a different electronegativity, the molecule becomes polar. A
trigonal planar molecule (BF ) may be nonpolar if all three peripheral atoms are the same, but a trigonal pyramidal molecule
3

(NH ) is polar.
3

Figure 2.23.8 : Some examples of polar and nonpolar molecules based on molecular geometry.
To summarize, to be polar, a molecule must:
1. Contain at least one polar covalent bond.
2. Have a molecular structure such that the sum of the vectors of each bond dipole moment do not cancel.

2.23.4 https://chem.libretexts.org/@go/page/423609
 Steps to Identify Polar Molecules
1. Draw the Lewis structure.
2. Figure out the geometry (using VSEPR theory).
3. Visualize or draw the geometry.
4. Find the net dipole moment (you don't have to actually do calculations if you can visualize it).
5. If the net dipole moment is zero, it is non-polar. Otherwise, it is polar.

Properties of Polar Molecules


Polar molecules tend to align when placed in an electric field with the positive end of the molecule oriented toward the negative
plate and the negative end toward the positive plate (Figure 2.23.14). We can use an electrically charged object to attract polar
molecules, but nonpolar molecules are not attracted. Also, polar solvents are better at dissolving polar substances, and nonpolar
solvents are better at dissolving nonpolar substances.

Figure 2.23.9 : (a) Molecules are always randomly distributed in the liquid state in the absence of an electric field. (b) When an
electric field is applied, polar molecules like HF will align to the dipoles with the field direction. (OpenStax CC-BY-SA);
While molecules can be described as "polar covalent" or "ionic", it must be noted that this is often a relative term, with one
molecule simply being more polar or less polar than another. However, the following properties are typical of such molecules.
Polar molecules tend to:
have higher melting points than nonpolar molecules
have higher boiling points than nonpolar molecules
be more soluble in water (dissolve better) than nonpolar molecules
have lower vapor pressures than nonpolar molecules

 Example 2.23.2:

Label each of the following as polar or nonpolar.

a. Water, H2O:

b. Methanol, CH3OH:
c. Hydrogen Cyanide, HCN:
d. Oxygen, O2:

e. Propane, C3H8:

2.23.5 https://chem.libretexts.org/@go/page/423609
Solution
a. Water is polar. Any molecule with lone pairs of electrons around the central atom is polar.
b. Methanol is polar. This is not a symmetric molecule. The −OH side is different from the other 3 −H sides.
c. Hydrogen cyanide is polar. The molecule is not symmetric. The nitrogen and hydrogen have different electronegativities,
creating an uneven pull on the electrons.
d. Oxygen is nonpolar. The molecule is symmetric. The two oxygen atoms pull on the electrons by exactly the same amount.
e. Propane is nonpolar, because it is symmetric, with H atoms bonded to every side around the central atoms and no unshared
pairs of electrons.

 Exercise 2.23.2

Label each of the following as polar or nonpolar.


a. SO 3

b. NH 3

Answer a
nonpolar

Answer b
polar

Contributions & Attributions


StackExchange (thomij).

2.23: Electronegativity and Polarity - Why Oil and Water Do not Mix is shared under a CK-12 license and was authored, remixed, and/or curated
by Marisa Alviar-Agnew & Henry Agnew.
10.8: Electronegativity and Polarity - Why Oil and Water Do not Mix by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12.
Original source: https://www.ck12.org/c/chemistry/.

2.23.6 https://chem.libretexts.org/@go/page/423609
2.24: A Molecular View of Elements and Compounds
 Learning Objectives
Classify substances as atomic elements, molecular elements, molecular compounds, or ionic compounds.

Atomic Elements
Most elements exist with individual atoms as their basic unit. It is assumed that there is only one atom in a formula if there is no
numerical subscript on the right side of an element’s symbol.

Molecular Elements
There are many substances that exist as two or more atoms connected together so strongly that they behave as a single particle.
These multi-atom combinations are called molecules. A molecule is the smallest part of a substance that has the physical and
chemical properties of that substance. In some respects, a molecule is similar to an atom. A molecule, however, is composed of
more than one atom.
Table 2.24.1 : Elements That Exist as Diatomic Molecules
Hydrogen, H Oxygen Nitrogen Fluorine Chlorine Bromine Iodine

Some elements exist naturally as molecules. For example, hydrogen and oxygen exist as two-atom molecules. Other elements also
exist naturally as diatomic molecules—a molecule with only two atoms (Table 2.24.1). As with any molecule, these elements are
labeled with a molecular formula, a formal listing of what and how many atoms are in a molecule. (Sometimes only the word
formula is used, and its meaning is inferred from the context.) For example, the molecular formula for elemental hydrogen is H2,
with H being the symbol for hydrogen and the subscript 2 implying that there are two atoms of this element in the molecule. Other
diatomic elements have similar formulas: O2, N2, and so forth. Other elements exist as molecules—for example, sulfur normally
exists as an eight-atom molecule, S8, while phosphorus exists as a four-atom molecule, P4 (Figure 2.24.1).

Figure 2.24.1 : Molecular Art of S8 and P4 Molecules. If each green ball represents a sulfur atom, then the diagram on the left
represents an S8 molecule. The molecule on the right shows that one form of elemental phosphorus exists, as a four-atom
molecule.
Figure 2.24.1 shows two examples of how molecules will be represented in this text. An atom is represented by a small ball or
sphere, which generally indicates where the nucleus is in the molecule. A cylindrical line connecting the balls represents the
connection between the atoms that make this collection of atoms a molecule. This connection is called a chemical bond.

Ionic Compounds
The elements in the periodic table are divided into specific groupings; the metals, the non-metals, the semi-metals, and so on. These
groupings are largely based on physical properties and on the tendency of the various elements to bond with other elements by
forming either an ionic or a covalent bond. As a general rule of thumb, compounds that involve a metal binding with either a non-
metal or a semi-metal will display ionic bonding. Thus, the compound formed from sodium and chlorine will be ionic (a metal and
a non-metal). The basic unit of ionic compounds is the formula unit.

Molecular Compounds
Compounds that are composed of only non-metals or semi-metals with non-metals will display covalent bonding and will be
classified as molecular compounds. Nitrogen monoxide (NO) will be a covalently bound molecule (two non-metals) and silicon

2.24.1 https://chem.libretexts.org/@go/page/423610
dioxide (SiO2) will also be a covalently bound molecule (a semi-metal and a non-metal). The basic unit of molecular compounds is
the molecule.

 Example 2.24.1

Provide the classification (i.e. atomic element, molecular element, molecular compound, or ionic compound) of each
substance.
a. Fe
b. PCl3
c. LiBr
d. P4
e. oxygen gas

Solution
a. Fe (iron) is an element that is represented with no subscript, so it is an atomic element.
b. PCl3 is made up of two nonmetals, so it is a molecular compound.
c. LiBr is made up of lithium, a metal, and bromine, a nonmetal, so it is an ionic compound.
d. P4 is a substance that is made up of four atoms of the same element, so it is a molecular element.
e. The formula for oxygen gas is O2 so it is a molecular element.

 Exercise 2.24.1

Provide the classification (i.e. atomic element, molecular element, molecular compound, or ionic compound) of each
substance.
a. I2
b. He
c. H2O
d. Al
e. CuCl

Answer a:
molecular element

Answer b:
atomic element

Answer c:
molecular compound

Answer d:
atomic element

Answer e:
ionic compound

2.24: A Molecular View of Elements and Compounds is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa
Alviar-Agnew & Henry Agnew.
5.4: A Molecular View of Elements and Compounds by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

2.24.2 https://chem.libretexts.org/@go/page/423610
2.25: Writing Formulas for Ionic Compounds
 Learning Objectives
Write the correct formula for an ionic compound.
Recognize polyatomic ions in chemical formulas.

Ionic compounds do not exist as molecules. In the solid state, ionic compounds are in crystal lattice containing many ions each of
the cation and anion. An ionic formula, like NaCl, is an empirical formula. This formula merely indicates that sodium chloride is
made of an equal number of sodium and chloride ions. Sodium sulfide, another ionic compound, has the formula Na S. This 2

formula indicates that this compound is made up of twice as many sodium ions as sulfide ions. This section will teach you how to
find the correct ratio of ions, so that you can write a correct formula.
If you know the name of a binary ionic compound, you can write its chemical formula. Start by writing the metal ion with its
charge, followed by the nonmetal ion with its charge. Because the overall compound must be electrically neutral, decide how many
of each ion is needed in order for the positive and negative charges to cancel each other out.

 Example 2.25.1: Aluminum Nitride and Lithium Oxide


Write the formulas for aluminum nitride and lithium oxide.

Solution
Solution to Example 5.5.1
Write the formula for aluminum nitride Write the formula for lithium oxide

1. Write the symbol and charge of the cation


(metal) first and the anion (nonmetal) Al
3 +
N
3 −
Li
+ 2 −
O

second.

total charge of cations = total charge of total charge of cations = total charge of
2. Use a multiplier to make the total charge anions anions
of the cations and anions equal to each other. 1(3+) = 1(3-) 2(1+) = 1(2-)
+3 = -3 +2 = -2

3. Use the multipliers as subscript for each


Al N Li O
ion. 1 1 2 1

4. Write the final formula. Leave out all


AlN Li O
charges and all subscripts that are 1. 2

An alternative way to writing a correct formula for an ionic compound is to use the crisscross method. In this method, the
numerical value of each of the ion charges is crossed over to become the subscript of the other ion. Signs of the charges are
dropped.

 Example 2.25.2: The Crisscross Method for Lead (IV) Oxide


Write the formula for lead (IV) oxide.

Solution
Solution to Example 5.5.2
Crisscross Method Write the formula for lead (IV) oxide

1. Write the symbol and charge of the cation (metal) first and the 4 + 2 −
Pb O
anion (nonmetal) second.

2.25.1 https://chem.libretexts.org/@go/page/423611
Crisscross Method Write the formula for lead (IV) oxide

2. Transpose only the number of the positive charge to become the


subscript of the anion and the number only of the negative charge to
become the subscript of the cation.

3. Reduce to the lowest ratio. Pb O


2 4

4. Write the final formula. Leave out all subscripts that are 1. PbO
2

 Exercise 2.25.2

Write the chemical formula for an ionic compound composed of each pair of ions.
a. the calcium ion and the oxygen ion
b. the 2+ copper ion and the sulfur ion
c. the 1+ copper ion and the sulfur ion

Answer a:
CaO
Answer b:
CuS
Answer c:
Cu2S

Be aware that ionic compounds are empirical formulas and so must be written as the
lowest ratio of the ions.

 Example 2.25.3: Sulfur Compound

Write the formula for sodium combined with sulfur.

Solution
Solution to Example 5.5.3
Crisscross Method Write the formula for sodium combined with sulfur

1. Write the symbol and charge of the cation (metal) first and the + 2 −
Na S
anion (nonmetal) second.

2. Transpose only the number of the positive charge to become the


subscript of the anion and the number only of the negative charge to
become the subscript of the cation.

3. Reduce to the lowest ratio. This step is not necessary.

4. Write the final formula. Leave out all subscripts that are 1. Na S
2

 Exercise 2.25.3
Write the formula for each ionic compound.
a. sodium bromide
b. lithium chloride
c. magnesium oxide

Answer a:

2.25.2 https://chem.libretexts.org/@go/page/423611
NaBr
Answer b:
LiCl
Answer c:
MgO

Polyatomic Ions
Some ions consist of groups of atoms bonded together and have an overall electric charge. Because these ions contain more than
one atom, they are called polyatomic ions. Polyatomic ions have characteristic formulas, names, and charges that should be
memorized. For example, NO3− is the nitrate ion; it has one nitrogen atom and three oxygen atoms and an overall 1− charge. Table
2.25.1 lists the most common polyatomic ions.

Table 2.25.1 : Some Polyatomic Ions


Name Formula

ammonium ion NH4+

acetate ion C2H3O2− (also written CH3CO2−)

carbonate ion CO32−

chromate ion CrO42−

dichromate ion Cr2O72−

hydrogen carbonate ion (bicarbonate ion) HCO3−

cyanide ion CN−

hydroxide ion OH−

nitrate ion NO3−

nitrite ion NO2−

permanganate ion MnO4−

phosphate ion PO43−

hydrogen phosphate ion HPO42−

dihydrogen phosphate ion H2PO4−

sulfate ion SO42−

hydrogen sulfate ion (bisulfate ion) HSO4−

sulfite ion SO32−

The rule for constructing formulas for ionic compounds containing polyatomic ions is the same as for formulas containing
monatomic (single-atom) ions: the positive and negative charges must balance. If more than one of a particular polyatomic ion is
needed to balance the charge, the entire formula for the polyatomic ion must be enclosed in parentheses, and the numerical
subscript is placed outside the parentheses. This is to show that the subscript applies to the entire polyatomic ion. An example is
Ba(NO3)2.

Writing Formulas for Ionic Compounds Containing Polyatomic Ions


Writing a formula for ionic compounds containing polyatomic ions also involves the same steps as for a binary ionic compound.
Write the symbol and charge of the cation followed by the symbol and charge of the anion.

2.25.3 https://chem.libretexts.org/@go/page/423611
 Example 2.25.4: Calcium Nitrate
Write the formula for calcium nitrate.

Solution
Solution to Example 5.5.4
Crisscross Method Write the formula for calcium nitrate

1. Write the symbol and charge of the cation (metal) first and the 2 + −
Ca NO
anion (nonmetal) second. 3

2. Transpose only the number of the positive charge to become the


subscript of the anion and the number only of the negative charge to
The 2+ charge on Ca becomes the subscript of NO3 and the 1- charge
become the subscript of the cation.
on NO3 becomes the subscript of Ca.

3. Reduce to the lowest ratio. Ca (NO )


1 3 2

4. Write the final formula. Leave out all subscripts that are 1. If there
Ca(NO )
is only 1 of the polyatomic ion, leave off parentheses. 3 2

 Example 2.25.5

Write the chemical formula for an ionic compound composed of the potassium ion and the sulfate ion.

Solution
Solution to Example 5.5.5
Explanation Answer

Potassium ions have a charge of 1+, while sulfate ions have a charge
of 2−. We will need two potassium ions to balance the charge on the K SO
2 4

sulfate ion, so the proper chemical formula is K SO .


2 4

 Exercise 2.25.5

Write the chemical formula for an ionic compound composed of each pair of ions.
a. the magnesium ion and the carbonate ion
b. the aluminum ion and the acetate ion

Answer a:
MgCO
3

Answer b:
Al(CH COO)
3 3

Recognizing Ionic Compounds


There are two ways to recognize ionic compounds.

Method 1
Compounds between metal and nonmetal elements are usually ionic. For example, CaBr contains a metallic element 2

(calcium, a group 2 [or 2A] metal) and a nonmetallic element (bromine, a group 17 [or 7A] nonmetal). Therefore, it is most
likely an ionic compound (in fact, it is ionic). In contrast, the compound NO contains two elements that are both nonmetals
2

(nitrogen, from group 15 [or 5A], and oxygen, from group 16 [or 6A]. It is not an ionic compound; it belongs to the category

2.25.4 https://chem.libretexts.org/@go/page/423611
of covalent compounds discussed elsewhere. Also note that this combination of nitrogen and oxygen has no electric charge
specified, so it is not the nitrite ion.

Method 2
Second, if you recognize the formula of a polyatomic ion in a compound, the compound is ionic. For example, if you see the
formula Ba(NO ) , you may recognize the “NO ” part as the nitrate ion, NO . (Remember that the convention for writing
3 2 3

3

formulas for ionic compounds is not to include the ionic charge.) This is a clue that the other part of the formula, Ba , is
actually the Ba ion, with the 2+ charge balancing the overall 2− charge from the two nitrate ions. Thus, this compound is
2 +

also ionic.

 Example 2.25.6

Identify each compound as ionic or not ionic.


a. Na O 2

b. PCl 3

c. NH Cl4

d. OF 2

Solution
Solution to Example 5.5.6
Explanation Answer

a. Sodium is a metal, and oxygen is a nonmetal. Therefore, Na O is


2
Na O , ionic
expected to be ionic via method 1. 2

b. Both phosphorus and chlorine are nonmetals. Therefore, PCl is 3


PCl , not ionic
not ionic via method 1 3

c. The NH in the formula represents the ammonium ion, NH , +


4 4
NH Cl , ionic
which indicates that this compound is ionic via method 2 4

d. Both oxygen and fluorine are nonmetals. Therefore, OF is not


2
OF , not ionic
ionic via method 1 2

 Exercise 2.25.6

Identify each compound as ionic or not ionic.


a. N O
2

b. FeCl 3

c. (NH ) 4 3
PO
4

d. SOCl 2

Answer a:
not ionic
Answer b:
ionic
Answer c:
ionic
Answer d:
not ionic

2.25.5 https://chem.libretexts.org/@go/page/423611
Summary
Formulas for ionic compounds contain the symbols and number of each atom present in a compound in the lowest whole number
ratio.

2.25: Writing Formulas for Ionic Compounds is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-
Agnew & Henry Agnew.
5.5: Writing Formulas for Ionic Compounds by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

2.25.6 https://chem.libretexts.org/@go/page/423611
2.26: Nomenclature- Naming Compounds
Nomenclature is the process of naming chemical compounds so that they can be easily identified as separate chemicals. The
primary function of chemical nomenclature is to ensure that a spoken or written chemical name leaves no ambiguity concerning
which chemical compound the name refers to—each chemical name should refer to a single substance. A less important aim is to
ensure that each substance has a single name, although a limited number of alternative names is acceptable in some cases.
Preferably, the name also conveys some information about the structure or chemistry of a compound. A common name will often
suffice to identify a chemical compound in a particular set of circumstances. To be more generally applicable, the name should
indicate at least the chemical formula. To be more specific still, the three-dimensional arrangement of the atoms may need to be
specified.

Contributions & Attributions


Wikipedia (CC-BY-SA-3.0)

2.26: Nomenclature- Naming Compounds is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew &
Henry Agnew.
5.6: Nomenclature- Naming Compounds by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

2.26.1 https://chem.libretexts.org/@go/page/423612
2.27: Naming Ionic Compounds
 Learning Objectives
To use the rules for naming ionic compounds.

After learning a few more details about the names of individual ions, you will be one step away from knowing how to name ionic
compounds. This section begins the formal study of nomenclature, the systematic naming of chemical compounds.

Naming Ions
The name of a monatomic cation is simply the name of the element followed by the word ion. Thus, Na+ is the sodium ion, Al3+ is
the aluminum ion, Ca2+ is the calcium ion, and so forth.
We have seen that some elements lose different numbers of electrons, producing ions of different charges (Figure 3.3). Iron, for
example, can form two cations, each of which, when combined with the same anion, makes a different compound with unique
physical and chemical properties. Thus, we need a different name for each iron ion to distinguish Fe2+ from Fe3+. The same issue
arises for other ions with more than one possible charge.
There are two ways to make this distinction. In the simpler, more modern approach, called the Stock system, an ion’s positive
charge is indicated by a roman numeral in parentheses after the element name, followed by the word ion. Thus, Fe2+ is called the
iron(II) ion, while Fe3+ is called the iron(III) ion. This system is used only for elements that form more than one common positive
ion. We do not call the Na+ ion the sodium(I) ion because (I) is unnecessary. Sodium forms only a 1+ ion, so there is no ambiguity
about the name sodium ion.
Table 2.27.1 : The Modern and Common System of Cation Names
Element Stem Charge Modern Name Common Name

2+ iron(II) ion ferrous ion


iron ferr-
3+ iron(III) ion ferric ion

1+ copper(I) ion cuprous ion


copper cupr-
2+ copper(II) ion cupric ion

2+ tin(II) ion stannous ion


tin stann-
4+ tin(IV) ion stannic ion

2+ lead(II) ion plumbous ion


lead plumb-
4+ lead(IV) ion plumbic ion

2+ chromium(II) ion chromous ion


chromium chrom-
3+ chromium(III) ion chromic ion

1+ gold(I) ion aurous ion


gold aur-
3+ gold(III) ion auric ion

The second system, called the common system, is not conventional but is still prevalent and used in the health sciences. This
system recognizes that many metals have two common cations. The common system uses two suffixes (-ic and -ous) that are
appended to the stem of the element name. The -ic suffix represents the greater of the two cation charges, and the -ous suffix
represents the lower one. In many cases, the stem of the element name comes from the Latin name of the element. Table 2.27.1
lists the elements that use the common system, along with their respective cation names.
Table 2.27.2 : Some Monatomic Anions
Ion Name

F− fluoride ion

Cl− chloride ion

2.27.1 https://chem.libretexts.org/@go/page/423613
Ion Name

Br− bromide ion

I− iodide ion

O2− oxide ion

S2− sulfide ion

P3− phosphide ion

N3− nitride ion

The name of a monatomic anion consists of the stem of the element name, the suffix -ide, and then the word ion. Thus, as we have
already seen, Cl− is “chlor-” + “-ide ion,” or the chloride ion. Similarly, O2− is the oxide ion, Se2− is the selenide ion, and so forth.
Table 2.27.2 lists the names of some common monatomic ions. The polyatomic ions have their own characteristic names, as
discussed earlier.

 Example 2.27.1

Name each ion.


a. Ca2+
b. S2−
c. SO32−
d. NH4+
e. Cu+

Solution
a. the calcium ion
b. the sulfide ion
c. the sulfite ion
d. the ammonium ion
e. the copper(I) ion or the cuprous ion

 Exercise 2.27.1

Name each ion.


a. Fe2+
b. Fe3+
c. SO42−
d. Ba2+
e. HCO3−

Answer a:
iron(II) ion
Answer b:
iron(III) ion
Answer c:
sulfate ion
Answer d:
barium ion
Answer e:
hydrogen carbonate ion or bicarbonate ion

2.27.2 https://chem.libretexts.org/@go/page/423613
 Example 2.27.2

Write the formula for each ion.


a. the bromide ion
b. the phosphate ion
c. the cupric ion
d. the magnesium ion

Solution
a. Br−
b. PO43−
c. Cu2+
d. Mg2+

 Exercise 2.27.2

Write the formula for each ion.


a. the fluoride ion
b. the carbonate ion
c. the stannous ion
d. the potassium ion

Answer a:
F-
Answer b:
CO32-
Answer c:
Sn 2+
Answer d:
K+

Naming Binary Ionic Compounds with a Metal that Forms Only One Type of Cation
A binary ionic compound is a compound composed of a monatomic metal cation and a monatomic nonmetal anion. The metal
cation is named first, followed by the nonmetal anion as illustrated in Figure 2.27.1 for the compound BaCl2. The word ion is
dropped from both parts.

2.27.3 https://chem.libretexts.org/@go/page/423613
Figure 2.27.1 : Naming BaC l2

Naming formula: Name of metal cation + base


name of nonmetal anion + suffix -ide. BaCl2 is named as barium chloride.
Subscripts in the formula do not affect the name.

 Example 2.27.3: Naming Ionic Compounds

Name each ionic compound.


a. CaCl2
b. AlF3
c. KCl

Solution
a. Using the names of the ions, this ionic compound is named calcium chloride.
b. The name of this ionic compound is aluminum fluoride.
c. The name of this ionic compound is potassium chloride

 Exercise 2.27.3

Name each ionic compound.


a. AgI
b. MgO
c. Ca3P2

Answer a:
silver iodide
Answer b:
magnesium oxide
Answer c:
calcium phosphide

Naming Binary Ionic Compounds with a Metal That Forms More Than One Type of Cation
If you are given a formula for an ionic compound whose cation can have more than one possible charge, you must first determine
the charge on the cation before identifying its correct name. For example, consider FeCl2 and FeCl3 . In the first compound, the iron

2.27.4 https://chem.libretexts.org/@go/page/423613
ion has a 2+ charge because there are two Cl− ions in the formula (1− charge on each chloride ion). In the second compound, the
iron ion has a 3+ charge, as indicated by the three Cl− ions in the formula. These are two different compounds that need two
different names. By the Stock system, the names are iron(II) chloride and iron(III) chloride (Figure 2.27.2).
Table 2.27.3 : Naming the F eC l and F eC l Compounds in the Modern/Stock System.
2 3

Name of cation (metal) + (Roman Numeral in parenthesis) + Base name of


anion (nonmetal) and -ide

If we were to use the stems and suffixes of the common system, the names would be ferrous chloride and ferric chloride,
respectively (Figure 2.27.3) .
Table 2.27.4 : Naming the F eC l and F eC l Compounds in the Old/Common System.
2 3

"Old" base name of cation (metal) and -ic or -ous + Base name of anion
(nonmetal) and -ide
-ous (for ions with lower charge) -ic (for ions with higher charge)

2.27.5 https://chem.libretexts.org/@go/page/423613
 Example 2.27.4:

Name each ionic compound.


a. Co2O3
b. FeCl2

Solution
Solutions to Example 5.7.4
Explanation Answer

We know that cobalt can have more than one


possible charge; we just need to determine
what it is.
Oxide always has a 2− charge, so with
three oxide ions, we have a total negative
a charge of 6−. cobalt(III) oxide
This means that the two cobalt ions have
to contribute 6+, which for two cobalt
ions means that each one is 3+.
Therefore, the proper name for this ionic
compound is cobalt(III) oxide.

Iron can also have more than one possible


charge.
Chloride always has a 1− charge, so with
two chloride ions, we have a total
b negative charge of 2−. iron(II) chloride
This means that the one iron ion must
have a 2+ charge.
Therefore, the proper name for this ionic
compound is iron(II) chloride.

 Exercise 2.27.4
Name each ionic compound.
a. AuCl3
b. PbO2
c. CuO

Answer a:
gold(III) chloride
Answer b:
lead(IV) oxide
Answer c:
copper(II) oxide

Naming Ionic Compounds with Polyatomic Ions


The process of naming ionic compounds with polyatomic ions is the same as naming binary ionic compounds. The cation is named
first, followed by the anion. One example is the ammonium sulfate compound in Figure 2.27.6.

2.27.6 https://chem.libretexts.org/@go/page/423613
Figure 2.27.2 : Naming Ionic Compounds with Polyatomic Ions

 Example 2.27.5: Naming Ionic Compounds

Write the proper name for each ionic compound.


a. (NH4)2S
b. AlPO4,
c. Fe3(PO4)2

Solution
Solutions to Example 5.7.5
Explanation Answer

a. The ammonium ion has a 1+ charge and the sulfide ion has a 2−
charge.
Two ammonium ions need to balance the charge on a single sulfide ammonium sulfide
ion.
The compound’s name is ammonium sulfide.

b. The ions have the same magnitude of charge, one of each (ion) is
needed to balance the charges. aluminum phosphate
The name of the compound is aluminum phosphate.

c. Neither charge is an exact multiple of the other, so we have to go


to the least common multiple of 6.
To get 6+, three iron(II) ions are needed, and to get 6−, two iron(II) phosphate
phosphate ions are needed .
The compound’s name is iron(II) phosphate.

 Exercise 2.27.5A
Write the proper name for each ionic compound.
a. (NH4)3PO4
b. Co(NO2)3

Answer a:
ammonium phosphate
Answer b:
cobalt(III) nitrite

2.27.7 https://chem.libretexts.org/@go/page/423613
Figure 2.27.1 is a synopsis of how to name simple ionic compounds.

Figure 2.27.3 : A Guide to Naming Simple Ionic Compounds.

 Exercise 2.27.5B
Name each ionic compound.
a. ZnBr2
b. Al2O3
c. (NH4)3PO4
d. AuF3
e. AgF

Answer a:
zinc bromide

Answer b:
aluminum oxide

Answer c:
ammonium phosphate

Answer d:
gold(III) fluoride or auric fluoride

Answer e:
silver fluoride

Summary
Ionic compounds are named by stating the cation first, followed by the anion.
Positive and negative charges must balance.
Some anions have multiple forms and are named accordingly with the use of roman numerals in parentheses.
Ternary compounds are composed of three or more elements.

2.27: Naming Ionic Compounds is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew & Henry
Agnew.
5.7: Naming Ionic Compounds by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

2.27.8 https://chem.libretexts.org/@go/page/423613
2.28: Naming Molecular Compounds
 Learning Objectives
Determine the name of a simple molecular compound from its chemical formula.

Molecular Compounds
Molecular compounds are inorganic compounds that take the form of discrete molecules. Examples include such familiar
substances as water (H O) and carbon dioxide (CO ). These compounds are very different from ionic compounds like sodium
2 2

chloride (NaCl). Ionic compounds are formed when metal atoms lose one or more of their electrons to nonmetal atoms. The
resulting cations and anions are electrostatically attracted to each other.
So what holds the atoms of a molecule together? Rather than forming ions, the atoms of a molecule share their electrons in such a
way that a bond forms between a pair of atoms. In a carbon dioxide molecule, there are two of these bonds, each occurring
between the carbon atom and one of the two oxygen atoms.

Figure 2.28.1 : Carbon dioxide molecules consist of a central carbon atom bonded to 2 oxygen atoms.
Larger molecules can have many, many bonds that serve to keep the molecule together. In a large sample of a given molecular
compound, all of the individual molecules are identical.

Naming Binary Molecular Compounds


Recall that a molecular formula shows the number of atoms of each element that a molecule contains. A molecule of water contains
two hydrogen atoms and one oxygen atom, so its formula is H O . A molecule of octane, which is a component of gasoline,
2

contains 8 atoms of carbon and 18 atoms of hydrogen. The molecular formula of octane is C H . 8 18

Figure 2.28.2 : Nitrogen dioxide ( NO )


2
is a reddish-brown toxic gas that is a prominent air pollutant produced by internal
combustion engines.
Naming binary (two-element) molecular compounds is similar to naming simple ionic compounds. The first element in the formula
is simply listed using the name of the element. The second element is named by taking the stem of the element name and adding
the suffix -ide. A system of numerical prefixes is used to specify the number of atoms in a molecule. Table 2.28.1 lists these
numerical prefixes.
Table 2.28.1 : Numerical Prefixes for Naming Binary Covalent Compounds
Number of Atoms in Compound Prefix on the Name of the Element

1 mono-*

2 di-

3 tri-

4 tetra-

5 penta-

6 hexa-

2.28.1 https://chem.libretexts.org/@go/page/423614
Number of Atoms in Compound Prefix on the Name of the Element

7 hepta-

8 octa-

9 nona-

10 deca-

*This prefix is not used for the first element’s name.

 Note
Generally, the less electronegative element is written first in the formula, though there are a few exceptions. Carbon is
always first in a formula and hydrogen is after nitrogen in a formula such as NH . The order of common nonmetals in
3

binary compound formulas is C, P, N , H , S, I , Br , Cl, O, F .


The a or o at the end of a prefix is usually dropped from the name when the name of the element begins with a vowel. As
an example, four oxygen atoms, is tetroxide instead of tetraoxide.
The prefix is "mono" is not added to the first element’s name if there is only one atom of the first element in a molecule.

Some examples of molecular compounds are listed in Table 2.28.2.


Table 2.28.2
Formula Name

NO nitrogen monoxide

N O
2
dinitrogen monoxide

S Cl
2 2
disulfur dichloride

Cl O
2 7
dichlorine heptoxide

Notice that the mono- prefix is not used with the nitrogen in the first compound, but is used with the oxygen in both of the first two
examples. The S Cl emphasizes that the formulas for molecular compounds are not reduced to their lowest ratios. The o of the
2 2

mono- and the a of hepta- are dropped from the name when paired with oxide.

 Exercise 2.28.1

Write the name for each compound.


a. CF4
b. SeCl2
c. SO3

Answer a:
carbon tetrafluoride

2.28.2 https://chem.libretexts.org/@go/page/423614
Answer b:
selenium dichloride
Answer c:
sulfur trioxide

Simple molecular compounds with common names


For some simple covalent compounds, we use common names rather than systematic names. We have already encountered these
compounds, but we list them here explicitly:
H2O: water
NH3: ammonia
CH4: methane
H2O2: hydrogen peroxide
Methane is the simplest organic compound. Organic compounds are compounds with carbon atoms and are named by a separate
nomenclature system.

Some Compounds Have Both Covalent and Ionic Bonds


If you recall the introduction of polyatomic ions, you will remember that the bonds that hold the polyatomic ions together are
covalent bonds. Once the polyatomic ion is constructed with covalent bonds, it reacts with other substances as an ion. The bond
between a polyatomic ion and another ion will be ionic. An example of this type of situation is in the compound sodium nitrate.
Sodium nitrate is composed of a sodium ion and a nitrate ion. The nitrate ion is held together by covalent bonds and the nitrate ion
is attached to the sodium ion by an ionic bond.

Summary
A molecular compound is usually composed of two or more nonmetal elements.
Molecular compounds are named with the first element first and then the second element by using the stem of the element name
plus the suffix -ide. Numerical prefixes are used to specify the number of atoms in a molecule.

2.28: Naming Molecular Compounds is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew &
Henry Agnew.
5.8: Naming Molecular Compounds by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

2.28.3 https://chem.libretexts.org/@go/page/423614
CHAPTER OVERVIEW

3: Module 3 - Stoichiometry
3.1: Counting Nails by the Pound
3.2: Counting Atoms by the Gram
3.3: Counting Molecules by the Gram
3.4: Chemical Formulas as Conversion Factors
3.5: Mass Percent Composition of Compounds
3.6: Mass Percent Composition from a Chemical Formula
3.7: The Chemical Equation
3.8: How to Write Balanced Chemical Equations
3.9: Making Pancakes- Relationships Between Ingredients
3.10: Making Molecules- Mole-to-Mole Conversions
3.11: Making Molecules- Mass-to-Mass Conversions
3.12: Stoichiometry
3.13: Limiting Reactant and Theoretical Yield
3.14: Limiting Reactant, Theoretical Yield, and Percent Yield from Initial Masses of Reactants
3.15: Enthalpy Change is a Measure of the Heat Evolved or Absorbed

3: Module 3 - Stoichiometry is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.

1
3.1: COUNTING NAILS BY THE POUND

COUNTING BY WEIGHING AND AVOGADRO'S NUMBER


The size of molecule is so small that it is physically difficult, if not impossible, to directly count out molecules (Figure 3.1.1 ). However, we
can count them indirectly by using a common trick of "counting by weighing".

Figure 3.1.1: A comparison of the scales of various biological and technological objects. (CC BY-SA 3.0; Wikipedia)
Consider the example of counting nails in a big box at a hardware store. You need to estimate the number of nails in a box. The weight of an
empty box is 213 g and the weight of the box plus a bunch of big nails is 1340 g . Assume that we know that the weight of one big nail is
0.450 g. Hopefully it's not necessary to tear open the package and count the nails. We agree that

mass of big nails = 1340 g − 213 g = 1227 g

Therefore
1227 g
Number of big nails in box = = 2, 726.6 big nails = 2, 730 big nails. (3.1.1)
0.450 g/big nail

You have just counted the number of big nails in the box by weighing them (rather than by counting them individually).

Figure 3.1.2: Galvanized nails. Individually counting nails in a box would would require significant effort. Alternatively, we can count them
by weighing. (Public Domain; Wikipedia).
Now consider if the box of nails weighed the same, but the box were filled with small nails with an individual mass of 0.23 g/small nail

instead? You would do the same math, but use a different denominator in Equation 3.1.1 :
1227 g
Number of small nails in box = = 5, 334.7 small nails = 5, 335 small nails. (3.1.2)
0.230 g/small nail

The individual mass is the conversion factor used in the calculation and changes, based on the nature of the nail (big or small). Let's ask a
different question: how many dozens of nails are there in the same box of small nails described above?

3.1.1 https://chem.libretexts.org/@go/page/423616
If we know the information from Equation 3.1.2 , we can just use the conversion of how many nails are in a dozen:
5, 335 small nails
= 444.6 dozen small nails (3.1.3)
12 small nails/dozen

If we want to get this value from weighing, we use the "dozen mass" instead of individual mass:
12 × 0.23g = 2.76 g/dozen small nails. (3.1.4)

So following Equation 3.1.2 , we get:


1227 g
Number of dozens of small nails = = 444.6 dozen small nails (3.1.5)
2.76 g/dozen small nails

and this is the same result as Equation 3.1.3 . These calculations demonstrate the difference between individual mass (i.e., per individual)
and collective mass (e.g., per dozen or per gross). The collective mass of most importance to chemistry is molar mass (i.e., mass per mole or
mass per 6.022 × 10 ).
23

 AVOGADRO'S NUMBER

Avogadro's number is an accident of nature. It is the number of particles that delivers a mole of a substance. Avogadro's number =
6.022 × 10
23
. The reason why the value is an accident of nature is that the mole is tied to the gram mass unit. The gram is a convenient
mass unit because it matches human sizes. If we were a thousand times greater in size (like Paul Bunyan) we would find it handy to use
kilogram amounts. This means the kilogram mole would be convenient. The number of particles handled in a kilogram mole is 1000
times greater. The kilo Avogadro number for the count of particles in a kilomole is 6.022 × 10 . 26

If humans were tiny creatures (like Lilliputians) only 1/1000 our present size, milligrams would be more convenient. This means the
milligram mole would be more useful. The number of particles handled in a milligram mole (millimole) would be 1/1000 times smaller.
The milli Avogadro number for the count of particles in a millimole is 6.022 × 10 . 20

What do you think would happen to Avogadro's number if the American system was used and amounts were measured in pound moles?
Remember 1 pound = 454 grams. Avogadro's number would be larger by a factor of 454. A pound mole of hydrogen would weigh 1
pound, which is 454 grams. A gram mole of hydrogen weighs 1 gram and contains 6.022 × 10 H atoms. 23

MOLAR MASS FOR ELEMENTS


You are able to read the periodic table and determine the average atomic mass for an element like carbon. The average mass is 12.01 amu.
This mass is a ridiculously tiny number of grams. It is too small to handle normally. The molar mass of carbon is defined as the mass in
grams that is numerically equal to the average atomic weight. This means

1g/molecarbon = 12.01 g carbon

this is commonly written


1 mol carbon = 12.01 grams carbon.

This is the mass of carbon that contains 6.022 × 10 23


carbon atoms.
Avogadro's number is 6.022 × 10 23
particles.
This same process gives us the molar mass of any element. For example:
1 mol neon = 20.18 g neon N e

1 mol sodium = 22.99 g sodium N a

MOLAR MASS FOR COMPOUNDS

 EXAMPLE 3.1.1: MOLAR MASS OF WATER

The formulas for compounds are familiar to you. You know the formula for water is H 2
O . It should be reasonable that the weight of a
formula unit can be calculated by adding up the weights for the atoms in the formula.

Solution
The formula weight for water
weight from hydrogen + weight from oxygen

3.1.2 https://chem.libretexts.org/@go/page/423616
The formula weight for water
2 H atoms x 1.008 amu + 1 O atom x 16.00 amu = 18.016 amu
The molar mass for water
18.016 grams water or 18 grams to the nearest gram

 EXAMPLE 3.1.2: MOLAR MASS OF METHANE

The formula for methane, the major component in natural gas, is CH . 4

Solution
The formula weight for methane
weight from hydrogen + weight from carbon
The formula weight for methane
4 H atoms x 1.008 amu + 1 C atom x 12.01 amu = 16.04 amu
The molar mass for methane
16.04 grams per mole of methane

 EXAMPLE 3.1.3: MOLAR MASS OF ETHYL CHLORIDE

What is its molar mass for ethyl chloride CH 3


CH Cl
2
?

Solution
The formula weight
weight from hydrogen + weight from carbon + weight from chlorine
The formula weight
5 H atoms x 1.008 amu + 2 C atom x 12.01 amu + 35.5 amu = 64.5 amu
The molar mass for ethyl chloride
64.5 grams per mole of ethyl chloride

 EXAMPLE 3.1.4: MOLAR MASS OF SULFUR DIOXIDE

What is the molar mass for sulfur dioxide, SO 2


, a gas used in bleaching and disinfection processes?
(g)

Solution
Look up the atomic weight for each of the elements in the formula.
1 sulfur atom = 32.07 amu
1 oxygen atom = 16.00 amu
Count the atoms of each element in the formula unit.
one sulfur atom
two oxygen atoms
The formula weight
weight from sulfur + weight from oxygen
The formula weight
1 sulfur atom x (32. 07 amu ) + 2 oxygen atoms x (16.00 amu)
The formula weight
SO
2
= 32. 07 amu + 32.00 amu = 64.07 amu = 64 amu SO 2

3.1.3 https://chem.libretexts.org/@go/page/423616
The molar mass for SO is 2

64.07 grams of SO ; 1 mol SO = 64 grams per mole of SO


2 2 2

 EXERCISE 3.1.1

What is the formula weight and molar mass for alum, KAl(SO )
4 2
∙ 12 H O
2
?

Answer
1. Check the periodic table for the atomic masses for each atom in the formula.
2. Count the number of each type of atom in the formula.
3. Multiply the number of atoms by the atomic mass for each element.
4. Add up the masses for all of the elements.
Table 3.1.1: Masses of each element in alum, KAl(SO 4
)
2
∙ 12 H O
2

element average atomic mass number of atoms in formula rounded to nearest one unit for simplicity
potassium k 39.1 amu 1 39. amu
aluminum 26.98 amu 1 27. amu
sulfur 32.07 amu 2 64. amu
oxygen 16.00 amu 8 + 12 = 20 320. amu
hydrogen 1.008 amu 2 x 12 = 24 24. amu

Molar mass is 474 grams (add up the amu of each element to find the total of 474 amu). This is a mass in grams that is numerically
(474) the same as the formula weight.
1 mole alum KAl(SO 4
)
2
∙ 12 H O
2
= 474 grams alum KAl(SO 4
)
2
∙ 12 H O
2

This page titled 3.1: Counting Nails by the Pound is shared under a CC BY-SA 3.0 license and was authored, remixed, and/or curated by Delmar Larsen,
Marisa Alviar-Agnew, Henry Agnew, & Henry Agnew.
6.2: Counting Nails by the Pound by Delmar Larsen, Henry Agnew, Marisa Alviar-Agnew is licensed CC BY-SA 3.0.

3.1.4 https://chem.libretexts.org/@go/page/423616
3.2: Counting Atoms by the Gram
 Learning Objectives
Use Avogadro's number to convert to moles and vice versa given the number of particles of an element.
Use the molar mass to convert to grams and vice versa given the number of moles of an element.

When objects are very small, it is often inconvenient, inefficient, or even impossible to deal with the objects one at a time. For
these reasons, we often deal with very small objects in groups, and have even invented names for various numbers of objects. The
most common of these is "dozen" which refers to 12 objects. We frequently buy objects in groups of 12, like doughnuts or pencils.
Even smaller objects such as straight pins or staples are usually sold in boxes of 144, or a dozen dozen. A group of 144 is called a
"gross".
This problem of dealing with things that are too small to operate with as single items also occurs in chemistry. Atoms and
molecules are too small to see, let alone to count or measure. Chemists needed to select a group of atoms or molecules that would
be convenient to operate with.

Avogadro's Number and Mole


In chemistry, it is impossible to deal with a single atom or molecule because we can't see them, count them, or weigh them.
Chemists have selected a number of particles with which to work that is convenient. Since molecules are extremely small, you may
suspect this number is going to be very large, and you are right. The number of particles in this group is 6.02 × 10 particles and
23

the name of this group is the mole (the abbreviation for mole is mol). One mole of any object is 6.02 × 10 of those objects.
23

There is a particular reason that this number was chosen and this reason will become clear as we proceed.
When chemists are carrying out chemical reactions, it is important that the relationship between the numbers of particles of each
reactant is known. Any readily measurable mass of an element or compound contains an extraordinarily large number of atoms,
molecules, or ions, so an extremely large numerical unit is needed to count them. The mole is used for this purpose.
The mole (symbol: mol) is the base unit of amount of substance ("number of substance") in the International System of Units or
System International (SI), defined as exactly 6.02214076×1023 particles, e.g., atoms, molecules, ions or electrons. The current
definition was adopted in November 2018, revising its old definition based on the number of atoms in 12 grams of carbon-12 (12C)
(the isotope of carbon with relative atomic mass 12 Daltons, by definition). For most purposes, 6.022 × 1023 provides an adequate
number of significant figures. Just as 1 mole of atoms contains 6.022 × 1023 atoms, 1 mole of eggs contains 6.022 × 1023 eggs. This
number is called Avogadro’s number, after the 19th-century Italian scientist who first proposed a relationship between the volumes
of gases and the numbers of particles they contain.
It is not obvious why eggs come in dozens rather than 10s or 14s, or why a ream of paper contains 500 sheets rather than 400 or
600. The definition of a mole—that is, the decision to base it on 12 g of carbon-12—is also arbitrary. The important point is that 1
mole of carbon—or of anything else, whether atoms, compact discs, or houses—always has the same number of objects: 6.022 ×
1023.

Converting Between Number of Atoms to Moles and Vice Versa


We can use Avogadro's number as a conversion factor, or ratio, in dimensional analysis problems. If we are given the number of
atoms of an element X, we can convert it into moles by using the relationship
23
1 mol X = 6.022 × 10 X atoms.

 Example 3.2.1: Moles of Carbon

The element carbon exists in two primary forms: graphite and diamond. How many moles of carbon atoms is 4.72 × 10
24

atoms of carbon?

Solution
Solutions to Example 6.3.1

3.2.1 https://chem.libretexts.org/@go/page/423617
The element carbon exists in two primary forms: graphite and
Steps for Problem Solving diamond. How many moles of carbon atoms is 4.72 × 10 24

atoms of carbon?

Identify the "given" information and what the problem is asking you Given: 4.72 × 10 C atoms
24

to "find." Find: mol C

List other known quantities. 1 mol = 6.022 × 10


23
C atoms

Prepare a concept map and use the proper conversion factor.

1 mol C
24
Cancel units and calculate. 4.72 × 10 C atoms ×
23
= 7.84 mol C
6.02 × 10 C atoms

The given number of carbon atoms was greater than Avogadro's


number, so the number of moles of C atoms is greater than 1 mole.
Think about your result. Since Avogadro's number is a measured quantity with three
significant figures, the result of the calculation is rounded to three
significant figures.

Molar Mass
Molar mass is defined as the mass of one mole of representative particles of a substance. By looking at a periodic table, we can
conclude that the molar mass of the element lithium is 6.94g, the molar mass of zinc is 65.38g, and the molar mass of gold is
196.97g. Each of these quantities contains 6.022 × 10 atoms of that particular element. The units for molar mass are grams per
23

mole or g/mol. 1.00 mol of carbon-12 atoms has a mass of 12.0 g and contains 6.022 × 10 atoms. 1.00 mole of any element has
23

a mass numerically equal to its atomic mass in grams and contains 6.022 × 10 particles. The mass, in grams, of 1 mole of
23

particles of a substance is now called the molar mass (mass of 1.00 mole).

Converting Grams to Moles of an Element and Vice Versa


We can also convert back and forth between grams of an element and moles. The conversion factor for this is the molar mass of the
substance. The molar mass is the ratio giving the number of grams for each one mole of the substance. This ratio is easily found by
referring to the atomic mass of the element using the periodic table. This ratio has units of grams per mole or g/mol.
Conversions like this are possible for any substance, as long as the proper atomic mass, formula mass, or molar mass is known (or
can be determined) and expressed in grams per mole. Figure 6.4.1 illustrates what conversion factor is needed and two examples
are given below.

Figure 3.2.1 : A Simple Flowchart for Converting Between Mass and Moles of an Element.
Flowchart: to convert moles of Element A to mass of Element A, use g/mol, and to convert vice versa, use mol/g.

 Example 3.2.2: Chromium

Chromium metal is used for decorative electroplating of car bumpers and other surfaces. Find the mass of 0.560 moles of
chromium.

3.2.2 https://chem.libretexts.org/@go/page/423617
Solution
Solutions to Example 6.3.2
Chromium metal is used for decorative electroplating of car
Steps for Problem Solving bumpers and other surfaces. Find the mass of 0.560 moles of
chromium.

Identify the "given" information and what the problem is asking you Given: 0.560 mol Cr
to "find." Find: g Cr

List other known quantities. 1 mol Cr = 52.00g Cr

Prepare a concept map and use the proper conversion factor.

52.00 g Cr
Cancel units and calculate. 0.560 mol Cr × = 29.1 g Cr
1 mol Cr

Since the desired amount was slightly more than one half of a mole,
Think about your result. the mass should be slightly more than one half of the molar mass. The
answer has three significant figures because of the 0.560 mol

 Example 3.2.3: Silicon

How many moles are in 107.6g of Si?

Solution
Solutions to Example 6.3.3
Steps for Problem Solving How many moles are in 107.6g of Si.

Identify the "given" information and what the problem is asking you Given: 107.6g Si
to "find." Find: mol Si

List other known quantities. 1 mol Si = 28.09g Si

Prepare a concept map and use the proper conversion factor.

1 mol Si

Cancel units and calculate. 107.6 g Si × = 3.83 mol Si


28.09 g Si

Think about your result. Since 1 mol of Si is 28.09g, 107.6 should be about 4 moles.

 Exercise 3.2.1
a. How many moles are present in 100.0 g of Al?
b. What is the mass of 0.552 mol of Ag metal?

Answer a:
3.706 mol Al

Answer b:
59.5 g Ag

3.2.3 https://chem.libretexts.org/@go/page/423617
Summary
A mole is defined as exactly 6.02214076×1023 particles, e.g., atoms, molecules, ions or electrons.
There are 6.02214076 × 10 particles in 1.00 mole. This number is called Avogadro's number.
23

The molar mass of an element can be found by referring to the atomic mass on a periodic table with units of g/mol.
Using dimensional analysis, it is possible to convert between grams, moles, and the number of atoms or molecules.

Further Reading/Supplemental Links


learner.org/resources/series61.html - The learner.org website allows users to view streaming videos of the Annenberg series of
chemistry videos. You are required to register before you can watch the videos, but there is no charge. The website has one
video that relates to this lesson called The Mole.
Using Avogadro's law, the mass of a substance can be related to the number of particles contained in that mass. The Mole:
(www.learner.org/vod/vod_window.html?pid=803)
Vision Learning tutorial: The Mole http://visionlearning.com/library/mo...p?mid-53&1=&c3=

Contributions & Attributions

Wikipedia

3.2: Counting Atoms by the Gram is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew & Henry
Agnew.
6.3: Counting Atoms by the Gram by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

3.2.4 https://chem.libretexts.org/@go/page/423617
3.3: COUNTING MOLECULES BY THE GRAM
 LEARNING OBJECTIVES
Define molecular mass and formula mass.
Perform conversions between mass and moles of a compound.
Perform conversions between mass and number of particles.

MOLECULAR AND FORMULA MASSES


The molecular mass of a substance is the sum of the average masses of the atoms in one molecule of a substance. It is calculated by adding
together the atomic masses of the elements in the substance, each multiplied by its subscript (written or implied) in the molecular formula.
Because the units of atomic mass are atomic mass units, the units of molecular mass are also atomic mass units. The procedure for
calculating molecular masses is illustrated in Example 3.3.1 .

 EXAMPLE 3.3.1: ETHANOL


Calculate the molecular mass of ethanol, whose condensed structural formula is CH 3
CH OH
2
. Among its many uses, ethanol is a fuel
for internal combustion engines

Solution
Solutions to Example 6.4.1
Calculate the molecular mass of ethanol, whose condensed
Steps for Problem Solving
structural formula is CH CH OH
3 2

Given: Ethanol molecule (CH3CH2OH)


Identify the "given"information and what the problem is asking you to "find."
Find: molecular mass
The molecular formula of ethanol may be written in three
different ways:
CH3CH2OH (which illustrates the presence of an ethyl
group
Determine the number of atoms of each element in the molecule.
CH3CH2−, and an −OH group)
C2H5OH, and C2H6O;
All show that ethanol has two carbon atoms, six hydrogen
atoms, and one oxygen atom.
1 C atom = 12.011 amu
Obtain the atomic masses of each element from the periodic table and multiply the atomic mass of
1 H atom = 1.0079 amu
each element by the number of atoms of that element.
1 O atom = 15.9994 amu
2C: (2 atoms)(12.011amu/atom) = 24.022 amu
6H: (6 atoms)(1.0079amu/atom) = 6.0474amu
Add the masses together to obtain the molecular mass.
+1O: (1 atoms)(15.9994amu/atom) =15.9994amu
C2H6O : molecular mass of ethanol = 46.069amu

 EXERCISE 3.3.1: FREON

Calculate the molecular mass of trichlorofluoromethane, also known as Freon-11, which has a condensed structural formula of CCl 3
.
F

Until recently, it was used as a refrigerant. The structure of a molecule of Freon-11 is as follows:

Figure 3.3.1: Molecular structure of freon-11, CCl 3


F .

3.3.1 https://chem.libretexts.org/@go/page/423618
Answer
137.37 amu

Unlike molecules, which form covalent bonds, ionic compounds do not have a readily identifiable molecular unit. Therefore, for ionic
compounds, the formula mass (also called the empirical formula mass) of the compound is used instead of the molecular mass. The
formula mass is the sum of the atomic masses of all the elements in the empirical formula, each multiplied by its subscript (written or
implied). It is directly analogous to the molecular mass of a covalent compound. The units are atomic mass units.

Atomic mass, molecular mass, and formula mass all have the same units: atomic mass units.

 EXAMPLE 3.3.2: CALCIUM PHOSPHATE

Calculate the formula mass of Ca (PO )


3 4 2
, commonly called calcium phosphate. This compound is the principal source of calcium
found in bovine milk.

Solution
Solutions to Example 6.4.2
Steps for Problem Solving Calculate the formula mass of Ca 3
(PO )
4 2
, commonly called calcium phosphate.
Given: Calcium phosphate [Ca3(PO4)2] formula unit
Identify the "given" information and what the problem is asking you to "find."
Find: formula mass
The empirical formula—Ca3(PO4)2—indicates that the simplest electrically neutral
unit of calcium phosphate contains three Ca2+ ions and two PO43− ions.
Determine the number of atoms of each element in the molecule. The formula mass of this molecular unit is calculated by adding together the
atomic masses of three calcium atoms, two phosphorus atoms, and eight oxygen
atoms.

1 Ca atom = 40.078 amu


Obtain the atomic masses of each element from the periodic table and multiply
1 P atom = 30.973761 amu
the atomic mass of each element by the number of atoms of that element.
1 O atom = 15.9994 amu
3Ca: (3 atoms) (40.078 amu/atom)=120.234amu
2P: (2 atoms) (30.973761amu/atom)=61.947522amu
Add together the masses to give the formula mass. + 8O: (8 atoms)(15.9994amu/atom)=127.9952amu

Formula mass of Ca3(PO4)2=310.177amu

 EXERCISE 3.3.2: SILICON NITRIDE

Calculate the formula mass of Si N , commonly called silicon nitride. It is an extremely hard and inert material that is used to make
3 4

cutting tools for machining hard metal alloys.

Figure 3.3.2: Si 3
N
4
bearing parts. (Public Domain; David W. Richerson and Douglas W. Freitag; Oak Ridge National Laboratory).

Answer
140.29 amu

3.3.2 https://chem.libretexts.org/@go/page/423618
MOLAR MASS
The molar mass of a substance is defined as the mass in grams of 1 mole of that substance. One mole of isotopically pure carbon-12 has a
mass of 12 g. For an element, the molar mass is the mass of 1 mol of atoms of that element; for a covalent molecular compound, it is the
mass of 1 mol of molecules of that compound; for an ionic compound, it is the mass of 1 mol of formula units. That is, the molar mass of a
substance is the mass (in grams per mole) of 6.022 × 1023 atoms, molecules, or formula units of that substance. In each case, the number of
grams in 1 mol is the same as the number of atomic mass units that describe the atomic mass, the molecular mass, or the formula mass,
respectively.

The molar mass of any substance is its atomic mass, molecular mass, or formula mass in grams per
mole.

The periodic table lists the atomic mass of carbon as 12.011 amu; the average molar mass of carbon—the mass of 6.022 × 1023 carbon
atoms—is therefore 12.011 g/mol:
Table 3.3.1: Molar Mass of Select Substances
Substance (formula) Basic Unit Atomic, Molecular, or Formula Mass (amu) Molar Mass (g/mol)
carbon (C) atom 12.011 (atomic mass) 12.011
ethanol (C2H5OH) molecule 46.069 (molecular mass) 46.069
calcium phosphate [Ca3(PO4)2] formula unit 310.177 (formula mass) 310.177

CONVERTING BETWEEN GRAMS AND MOLES OF A COMPOUND


The molar mass of any substance is the mass in grams of one mole of representative particles of that substance. The representative particles
can be atoms, molecules, or formula units of ionic compounds. This relationship is frequently used in the laboratory. Suppose that for a
certain experiment you need 3.00 moles of calcium chloride (CaCl ) . Since calcium chloride is a solid, it would be convenient to use a
2

balance to measure the mass that is needed. Dimensional analysis will allow you to calculate the mass of CaCl that you should measure as
2

shown in Example 3.3.3 .

 EXAMPLE 3.3.3: CALCIUM CHLORIDE

Calculate the mass of 3.00 moles of calcium chloride (CaCl2).

Figure 3.3.3: Calcium chloride is used as a drying agent and as a road deicer.

Solution

3.3.3 https://chem.libretexts.org/@go/page/423618
Solutions to Example 6.4.3
Steps for Problem Solving Calculate the mass of 3.00 moles of calcium chloride
Given: 3.00 moles of CaCl
Identify the "given" information and what the problem is asking you to "find." 2

Find: g CaCl 2

List other known quantities. 1 mol CaCl = 110.98 g CaCl


2 2

Prepare a concept map and use the proper conversion factor.

110.98 g CaCl
2

Cancel units and calculate. 3.00 mol CaCl


2
× = 333 g CaCl
2
1 mol CaCl
2

Think about your result.

 EXERCISE 3.3.3: CALCIUM OXIDE

What is the mass of 7.50 mol of (calcium oxide) CaO?

Answer
420.60 g

 EXAMPLE 3.3.4: WATER

How many moles are present in 108 grams of water?

Solution
Solutions to Example 6.4.4
Steps for Problem Solving How many moles are present in 108 grams of water?
Given: 108 g H O
Identify the "given" information and what the problem is asking you to "find." 2

Find: mol H O 2

List other known quantities. 1 mol H O = 18.02 g


2
H2O

Prepare a concept map and use the proper conversion factor.

1 mol H O
2

Cancel units and calculate. 108 g H O ×


2
= 5.99 mol H O
2
18.02 g H O
2

Think about your result.

 EXERCISE 3.3.4: NITROGEN GAS


What is the mass of 7.50 mol of Nitrogen gas N ? 2

Answer
210 g

3.3.4 https://chem.libretexts.org/@go/page/423618
CONVERSIONS BETWEEN MASS AND NUMBER OF PARTICLES
In "Conversions Between Moles and Mass", you learned how to convert back and forth between moles and the number of representative
particles. Now you have seen how to convert back and forth between moles and mass of a substance in grams. We can combine the two
types of problems into one. Mass and number of particles are both related to moles. To convert from mass to number of particles or vice-
versa, it will first require a conversion to moles as shown in Figure 3.3.1 and Example 3.3.5 .

Figure 3.3.4: Conversion from number of particles to mass, or from mass to number of particles, requires two steps. To convert from
number of particles to moles, use mol/Avogrado's #, an d to convert from moles to mass, use g/mol.

 EXAMPLE 3.3.5: CHLORINE

How many molecules is 20.0 g of chlorine gas, Cl ? 2

Solution
Solutions to Example 6.4.5
Steps for Problem Solving How many molecules is 20.0 g of chlorine gas, Cl ? 2

Identify the "given" information and what the problem is Given: 20.0 g Cl 2

asking you to "find." Find: # Cl molecules


2

1 mol Cl = 70.90 g Cl ,
List other known quantities.
2 2

1mol Cl = 6.022 × 10
2
23
Cl
2
molecules

Prepare a concept map and use the proper conversion


factor.

The conversion factors are 1 mole Cl


2
over 70.90 grams Cl
2
, and 2
6 . 0 2 2 × 1 0 3 Cl
2
molecules over 1
mole Cl .
2

1 mol Cl 23
2 6.02 × 10 molecules Cl
2
20.0 g Cl × ×
2
Cancel units and calculate. 70.90 g Cl
2
1 mol Cl
2

23
= 1.70 × 10 molecules Cl
2

Since the given mass is less than half of the molar mass of chlorine, the resulting number of molecules is
Think about your result.
less than half of Avogadro's number.

 EXERCISE 3.3.5: CALCIUM CHLORIDE

How many formula units are in 25.0 g of CaCl ? 2

Answer
1.36 x 1023 CaCl formula units
2

3.3.5 https://chem.libretexts.org/@go/page/423618
SUMMARY
Calculations for formula mass and molecular mass are described.
Calculations involving conversions between moles of a material and the mass of that material are described.
Calculations are illustrated for conversions between mass and number of particles.

This page titled 3.3: Counting Molecules by the Gram is shared under a CC BY-NC-SA 3.0 license and was authored, remixed, and/or curated by Marisa
Alviar-Agnew & Henry Agnew.
6.4: Counting Molecules by the Gram by Henry Agnew, Marisa Alviar-Agnew is licensed CC BY-NC-SA 3.0.

3.3.6 https://chem.libretexts.org/@go/page/423618
3.4: Chemical Formulas as Conversion Factors
 Learning Objectives
Use chemical formulas as conversion factors.

Figure 3.4.1 shows that we need 2 hydrogen atoms and 1 oxygen atom to make one water molecule. If we want to make two water
molecules, we will need 4 hydrogen atoms and 2 oxygen atoms. If we want to make five molecules of water, we need 10 hydrogen atoms
and 5 oxygen atoms. The ratio of atoms we will need to make any number of water molecules is the same: 2 hydrogen atoms to 1 oxygen
atom.

Figure 3.4.1 Water Molecules. The ratio of hydrogen atoms to oxygen atoms used to make water molecules is always 2:1, no matter how
many water molecules are being made.
Using formulas to indicate how many atoms of each element we have in a substance, we can relate the number of moles of molecules to
the number of moles of atoms. For example, in 1 mol of water (H2O) we can construct the relationships given in (Table 3.4.1).
Table 3.4.1 : Molecular Relationships for Water
1 Molecule of H 2O Has 1 Mol of H 2O Has Molecular Relationships
2 mol H atoms
or
1 mol H2 O molecules
2 H atoms 2 mol of H atoms
1 mol H2 O molecules

2 mol H atoms

1 mol O atoms
or
1 mol H2 O molecules
1 O atom 1 mol of O atoms
1 mol H2 O molecules

1 mol O atoms

 The Mole is big


A mole represents a very large number! The number 602,214,129,000,000,000,000,000 looks about twice as long as a trillion, which
means it’s about a trillion trillion.

(CC BY-SA NC; what if? [what-if.xkcd.com]).


A trillion trillion kilograms is how much a planet weighs. If 1 mol of quarters were stacked in a column, it could stretch back and
forth between Earth and the sun 6.8 billion times.

Table 3.4.2 : Molecular and Mass Relationships for Ethanol


1 Molecule of C2 H6 O Has 1 Mol of C 2 H6 O Has Molecular and Mass Relationships
2 mol C atoms
or
1 mol C2 H6 O molecules
2 C atoms 2 mol of C atoms
1 mol C2 H6 O molecules

2 mol C atoms

6 mol H atoms
or
1 mol C2 H6 O molecules
6 H atoms 6 mol of H atoms
1 mol C2 H6 O molecules

6 mol H atoms

3.4.1 https://chem.libretexts.org/@go/page/423619
1 Molecule of C 2 H6 O Has 1 Mol of C 2 H6 O Has Molecular and Mass Relationships
1 mol O atoms
or
1 mol C2 H6 O molecules
1 O atom 1 mol of O atoms
1 mol C2 H6 O molecules

1 mol O atoms

24.02 g C
or
2 (12.01 amu) C 2 (12.01 g) C 1 mol C2 H6 O molecules

24.02 amu C 24.02 g C 1 mol C2 H6 O molecules

24.02 g C

6.048 g H
or
6 (1.008 amu) H 6 (1.008 g) H 1 mol C2 H6 O molecules

6.048 amu H 6.048 g H 1 mol C2 H6 O molecules

6.048 g H

16.00 g O
or
1 (16.00 amu) O 1 (16.00 g) O 1 mol C2 H6 O molecules

16.00 amu O 16.00 g O 1 mol C2 H6 O molecules

16.00 g O

The following example illustrates how we can use the relationships in Table 3.4.2 as conversion factors.

 Example 3.4.1: Ethanol

If a sample consists of 2.5 mol of ethanol (C2H6O), how many moles of carbon atoms does it have?

Solution
Solutions to Example 6.5.1
If a sample consists of 2.5 mol of ethanol (C2H6O), how many moles
Steps for Problem Solving
of carbon atoms does it have?

Identify the "given" information and what the problem is asking you to Given: 2.5 mol C2H6O
"find." Find: mol C atoms

List other known quantities. 1 mol C2H6O = 2 mol C

Prepare a concept map and use the proper conversion factor.

Note how the unit mol C2H6O molecules cancels algebraically.


Cancel units and calculate. 2.5 mol C2 H6 O molecules ×
2 mol C atoms
= 5.0 mol C atoms
1 mol C2 H6 O molecules

There are twice as many C atoms in one C2H6O molecule, so the final
Think about your result.
amount should be double.

 Exercise 3.4.1
If a sample contains 6.75 mol of Na2SO4, how many moles of sodium atoms, sulfur atoms, and oxygen atoms does it have?

Answer
13.5 mol Na atoms, 6.75 mol S atoms, and 27.0 mol O atoms

The fact that 1 mol equals 6.022 × 1023 items can also be used as a conversion factor.

3.4.2 https://chem.libretexts.org/@go/page/423619
 Example 3.4.2: Oxygen Mass

Determine the mass of Oxygen in 75.0g of C2H6O.

Solution
Solutions to Example 6.5.2
Steps for Problem Solving Determine the mass of Oxygen in 75.0g of C2H6O

Identify the "given" information and what the problem is asking you to Given: 75.0g C2H6O
"find." Find: g O

1 mol O = 16.0g O
List other known quantities. 1 mol C2H6O = 1 mol O
1 mol C2H6O = 46.07g C2H6O

Prepare a concept map and use the proper conversion factor.

The conversion factors are 1 mol C2H6O over 46.07 g C2H6O, 1 mol O
over 1 mol C2H6O, and 16.00 g O over 1 mole O.

1 mol C2 H6 O 1 mol O 16.00 g O


Cancel units and calculate. 75.0 g C2 H6 O × × × = 26.0 g
46.07 g C2 H6 O 1 mol C2 H6 O 1 mol O

Think about your result.

 Exercise 3.4.2
a. How many molecules are present in 16.02 mol of C4H10? How many C atoms are in 16.02 mol?
b. How many moles of each type of atom are in 2.58 mol of Na2SO4?

Answer a:
9.647 x 1024 C4H10 molecules and 3.859 x 1025 C atoms

Answer b:
5.16 mol Na atoms, 2.58 mol S atoms, and 10.3 mol O atoms

Summary
In any given formula, the ratio of the number of moles of molecules (or formula units) to the number of moles of atoms can be used as a
conversion factor.

3.4: Chemical Formulas as Conversion Factors is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew &
Henry Agnew.
6.5: Chemical Formulas as Conversion Factors by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

3.4.3 https://chem.libretexts.org/@go/page/423619
3.5: Mass Percent Composition of Compounds
 Learning Objectives
Determine percent composition of each element in a compound based on mass.

Packaged foods that you eat typically have nutritional information provided on the label. The label on a jar of peanut butter reveals
that one serving size is considered to be 32 g. The label also gives the masses of various types of compounds that are present in
each serving. One serving contains 7 g of protein, 15 g of fat, and 3 g of sugar. By calculating the fraction of protein, fat, or sugar
in one serving size of peanut butter and converting to percent values, we can determine the composition of peanut butter on a
percent by mass basis.

Percent Composition
Chemists often need to know what elements are present in a compound and in what percentage. The percent composition is the
percent by mass of each element in a compound. It is calculated in a similar way to that of the composition of the peanut butter.
mass of element
% by mass = × 100%
mass of compound

The sample problem below shows the calculation of the percent composition of a compound based on mass data.

 Example 3.5.1: Percent Composition from Mass Data


A certain newly synthesized compound is known to contain the elements zinc and oxygen. When a 20.00 g sample of the
compound is decomposed, 16.07 g of zinc remains. Determine the percent composition of the compound.

Solution
Solutions to Example 6.6.1
When a 20.00 g sample of the zinc-and-oxygen compound is
Steps for Problem Solving decomposed, 16.07 g of zinc remains. Determine the percent
composition of the compound.

Given : Mass of compound = 20.00 g


Identify the "given" information and what the problem is asking you
Mass of Zn = 16.07 g
to "find."
Find: % Composition (% Zn and %O)
Subtract to find the mass of oxygen in the compound. Divide each
element's mass by the mass of the compound to find the percent by
List other known quantities.
mass.
Mass of oxygen = 20.00 g - 16.07 g = 3.93 g O

16.07 g Zn
% Zn = × 100% = 80.35% Zn
20.00 g

3.93 g O
%O = × 100% = 19.65% O
Cancel units and calculate. 20.00 g

Calculate the percent by mass of each element by dividing the mass


of that element by the mass of the compound and multiplying by
100%.

The calculations make sense because the sum of the two percentages
Think about your result.
adds up to 100%. By mass, the compound is mostly zinc.

3.5.1 https://chem.libretexts.org/@go/page/423620
 Exercise 3.5.1

Sulfuric acid, H2SO4 is a very useful chemical in industrial processes. If 196.0 g of sulfuric acid contained 64.0g oxygen and
4.0 g of hydrogen, what is the percent composition of the compound?

Answer
2.04% H, 32.65% S, and 65.3% O

Summary
Processes are described for calculating the percent composition of a compound based on mass.

3.5: Mass Percent Composition of Compounds is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-
Agnew & Henry Agnew.
6.6: Mass Percent Composition of Compounds by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

3.5.2 https://chem.libretexts.org/@go/page/423620
3.6: Mass Percent Composition from a Chemical Formula
 Learning Objectives
Determine the percent composition of each element in a compound from the chemical formula.

The percent composition of a compound can also be determined from the formula of the compound. The subscripts in the formula
are first used to calculate the mass of each element in one mole of the compound. This is divided by the molar mass of the
compound and multiplied by 100%.
mass of element in 1 mol
% by mass = × 100%
molar mass of compound

The percent composition of a given compound is always the same, given that the compound is pure.

 Example 3.6.1
Dichlorine heptoxide (Cl O ) is a highly reactive compound used in some organic synthesis reactions. Calculate the percent
2 7

composition of dichlorine heptoxide.

Solution
Solutions to Example 6.7.1
Calculate the percent composition of dichlorine heptoxide
Steps for Problem Solving
(Cl O ) .
2 7

Identify the "given" information and what the problem is asking you Given : Cl2O7
to "find." Find: % Composition (% Cl and %O)

Mass of Cl in 1 mol Cl2O7 , 2 Cl : 2 x 35.45 g = 70.90 g


List other known quantities. Mass of O in 1 mol Cl2O7 , 7 O: 7 x 16.00 g = 112.00 g
Molar mass of Cl2O7 = 182.90 g/mol

70.90 g Cl
%Cl = × 100% = 38.76% Cl
182.90 g

112.00 g O
%O = × 100% = 61.24% O
Cancel units and calculate. 182.90 g

Calculate the percent by mass of each element by dividing the mass


of that element in 1 mole of the compound by the molar mass of the
compound and multiplying by 100%.

Think about your result. The percentages add up to 100%.

Percent composition can also be used to determine the mass of a certain element that is contained in any mass of a compound. In
the previous sample problem, it was found that the percent composition of dichlorine heptoxide is 38.76% Cl and 61.24% O.
Suppose that you needed to know the masses of chlorine and oxygen present in a 12.50 g sample of dichlorine heptoxide. You can
set up a conversion factor based on the percent by mass of each element.
38.76 g Cl
12.50 g Cl O × = 4.845 g Cl
2 7
100 g Cl O
2 7

61.24 g O
12.50 g Cl O × = 7.655 g O
2 7
100 g Cl O
2 7

The sum of the two masses is 12.50 g, the mass of the sample size.

3.6.1 https://chem.libretexts.org/@go/page/423621
 Exercise 3.6.1

Barium fluoride is a transparent crystal that can be found in nature as the mineral frankdicksonite. Determine the percent
composition of barium fluoride.

Answer a:
78.32% Ba and 21.67% F

Summary
Processes are described for calculating the percent composition of a compound based on the chemical formula.

3.6: Mass Percent Composition from a Chemical Formula is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa
Alviar-Agnew & Henry Agnew.
6.7: Mass Percent Composition from a Chemical Formula by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

3.6.2 https://chem.libretexts.org/@go/page/423621
3.7: The Chemical Equation
 Learning Objectives
Identify the reactants and products in any chemical reaction.
Convert word equations into chemical equations.
Use the common symbols, (s) , (l), (g) , (aq) , and → appropriately when writing a chemical reaction.

In a chemical change, new substances are formed. In order for this to occur, the chemical bonds of the substances break, and the
atoms that compose them separate and rearrange themselves into new substances with new chemical bonds. When this process
occurs, we call it a chemical reaction. A chemical reaction is the process in which one or more substances are changed into one or
more new substances.

Reactants and Products


To describe a chemical reaction, we need to indicate what substances are present at the beginning and what substances are present
at the end. The substances that are present at the beginning are called reactants and the substances present at the end are called
products.
Sometimes when reactants are put into a reaction vessel, a reaction will take place to produce products. Reactants are the starting
materials, that is, whatever we have as our initial ingredients. The products are just that—what is produced—or the result of what
happens to the reactants when we put them together in the reaction vessel. If we think about baking chocolate chip cookies, our
reactants would be flour, butter, sugar, vanilla, baking soda, salt, egg, and chocolate chips. What would be the products? Cookies!
The reaction vessel would be our mixing bowl.

Flour + Butter + Sugar + Vanilla + Baking Soda + Eggs + Chocolate Chips → Cookies
 
Product
Ingredients = Reactants

Writing Chemical Equations


When sulfur dioxide is added to oxygen, sulfur trioxide is produced. Sulfur dioxide and oxygen, SO
2
+O
2
, are reactants and
sulfur trioxide, SO , is the product.
3

2 SO (g) + O (g) → 2 SO (g)


2 2 3
 
Reactants Products

In chemical reactions, the reactants are found before the symbol "→" and the products are found after the symbol "→". The general
equation for a reaction is:

Reactants → Products

There are a few special symbols that we need to know in order to "talk" in chemical shorthand. In the table below is the summary
of the major symbols used in chemical equations. Table 3.7.1 shows a listing of symbols used in chemical equations.
Table 3.7.1 : Symbols Used in Chemical Equations
Symbol Description Symbol Description

used to separate multiple reactants reactant or product in the solid


+ (s)
or products state

yield sign; separates reactants (l) reactant or product in the liquid



from products state

replaces the yield sign for


⇌ reversible reactions that reach (g) reactant or product in the gas state
equilibrium

Pt formula written above the arrow is reactant or product in an aqueous


(aq)

used as a catalyst in the reaction solution (dissolved in water)

3.7.1 https://chem.libretexts.org/@go/page/423622
Symbol Description Symbol Description

Δ triangle indicates that the reaction



is being heated

Chemists have a choice of methods for describing a chemical reaction.


1. They could draw a picture of the chemical reaction.

2. They could write a word equation for the chemical reaction:


"Two molecules of hydrogen gas react with one molecule of oxygen gas to produce two molecules of water vapor."
3. They could write the equation in chemical shorthand.

2H (g) + O (g) → 2 H O (g)


2 2 2

In the symbolic equation, chemical formulas are used instead of chemical names for reactants and products, while symbols are used
to indicate the phase of each substance. It should be apparent that the chemical shorthand method is the quickest and clearest
method for writing chemical equations.
We could write that an aqueous solution of calcium nitrate is added to an aqueous solution of sodium hydroxide to produce solid
calcium hydroxide and an aqueous solution of sodium nitrate. Or in shorthand we could write:

Ca (NO ) (aq) + 2NaOH (aq) → Ca (OH) (s) + 2 NaNO (aq)


3 2 2 3

How much easier is that to read? Let's try it in reverse. Look at the following reaction in shorthand and write the word equation for
the reaction:

Cu (s) + AgNO (aq) → Cu(NO ) (aq) + Ag (s)


3 3 2

The word equation for this reaction might read something like "solid copper reacts with an aqueous solution of silver nitrate to
produce a solution of copper (II) nitrate with solid silver."
To turn word equations into symbolic equations, we need to follow the given steps:
1. Identify the reactants and products. This will help you know which symbols go on each side of the arrow and where the + signs
go.
2. Write the correct formulas for all compounds. You will need to use the rules you learned in Chapter 5 (including making all
ionic compounds charge balanced).
3. Write the correct formulas for all elements. Usually this is given straight off of the periodic table. However, there are seven
elements that are considered diatomic, meaning that they are always found in pairs in nature. They include those elements listed
in the table.
Table 3.7.1 : Diatomic Elements
Element Name Hydrogen Nitrogen Oxygen Fluorine Chlorine Bromine Iodine

Formula H2 N2 O2 F2 Cl2 Br2 I2

 Example 3.7.1

Transfer the following symbolic equations into word equations or word equations into symbolic equations.
a. HCl (aq) + NaOH (aq) → NaCl (aq) + H O (l) 2

b. Gaseous propane, C H , burns in oxygen gas to produce gaseous carbon dioxide and liquid water.
3 8

c. Hydrogen fluoride gas reacts with an aqueous solution of potassium carbonate to produce an aqueous solution of potassium
fluoride, liquid water, and gaseous carbon dioxide.

3.7.2 https://chem.libretexts.org/@go/page/423622
Solution
a. An aqueous solution of hydrochloric acid reacts with an aqueous solution of sodium hydroxide to produce an aqueous
solution of sodium chloride and liquid water.
b. Reactants: propane (C 3
H
8
) and oxygen (O )2

Product: carbon dioxide (CO ) and water (H


2 2
O )

C H (g) + O (g) → CO (g) + H O (l)


3 8 2 2 2

c. Reactants: hydrogen fluoride and potassium carbonate


Products: potassium fluoride, water, and carbon dioxide

HF (g) + K CO (aq) → KF (aq) + H O (l) + CO (g)


2 3 2 2

 Exercise 3.7.1

Transfer the following symbolic equations into word equations or word equations into symbolic equations.
a. Hydrogen gas reacts with nitrogen gas to produce gaseous ammonia.
b. HCl (aq) + LiOH (aq) → LiCl (aq) + H O (l) 2

c. Copper metal is heated with oxygen gas to produce solid copper(II) oxide.

Answer a
H2 (g) + N2 (g) → N H3 (g)

Answer b
An aqueous solution of hydrochloric acid reacts with an aqueous solution of lithium hydroxide to produce an aqueous
solution of lithium chloride and liquid water.
Answer c
C u(s) + O2 (g) → C uO(s)

Summary
A chemical reaction is the process by which one or more substances are changed into one or more new substances.
Chemical reactions are represented by chemical equations.
Chemical equations have reactants on the left, an arrow that is read as "yields", and the products on the right.

3.7: The Chemical Equation is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew & Henry
Agnew.
7.3: The Chemical Equation by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

3.7.3 https://chem.libretexts.org/@go/page/423622
3.8: How to Write Balanced Chemical Equations
 Learning Objectives
Explain the roles of subscripts and coefficients in chemical equations.
Balance a chemical equation when given the unbalanced equation.
Explain the role of the Law of Conservation of Mass in a chemical reaction.

Even though chemical compounds are broken up and new compounds are formed during a chemical reaction, atoms in the reactants
do not disappear, nor do new atoms appear to form the products. In chemical reactions, atoms are never created or destroyed. The
same atoms that were present in the reactants are present in the products—they are merely reorganized into different arrangements.
In a complete chemical equation, the two sides of the equation must be present on the reactant and the product sides of the
equation.

Coefficients and Subscripts


There are two types of numbers that appear in chemical equations. There are subscripts, which are part of the chemical formulas of
the reactants and products; and there are coefficients that are placed in front of the formulas to indicate how many molecules of that
substance is used or produced.

Figure 3.8.1 : Balancing Equations. You cannot change subscripts in a chemical formula to balance a chemical equation; you can
change only the coefficients. Changing subscripts changes the ratios of atoms in the molecule and the resulting chemical
properties. For example, water (H2O) and hydrogen peroxide (H2O2) are chemically distinct substances. H2O2 decomposes to H2O
and O2 gas when it comes in contact with the metal platinum, whereas no such reaction occurs between water and platinum.
The subscripts are part of the formulas and once the formulas for the reactants and products are determined, the subscripts may not
be changed. The coefficients indicate the number of each substance involved in the reaction and may be changed in order to
balance the equation. The equation above indicates that one mole of solid copper is reacting with two moles of aqueous silver
nitrate to produce one mole of aqueous copper (II) nitrate and two atoms of solid silver.

Balancing a Chemical Equation


Because the identities of the reactants and products are fixed, the equation cannot be balanced by changing the subscripts of the
reactants or the products. To do so would change the chemical identity of the species being described, as illustrated in Figure 3.8.1.

Original molecule H2O: if the coefficient 2 is added in front, that makes 2 water molecules; but if the subscript 2 is added to make
H2O2, that's hydrogen peroxide.
The simplest and most generally useful method for balancing chemical equations is “inspection,” better known as trial and error.
The following is an efficient approach to balancing a chemical equation using this method.

3.8.1 https://chem.libretexts.org/@go/page/423623
 Steps in Balancing a Chemical Equation
1. Identify the most complex substance.
2. Beginning with that substance, choose an element(s) that appears in only one reactant and one product, if possible. Adjust
the coefficients to obtain the same number of atoms of this element(s) on both sides.
3. Balance polyatomic ions (if present on both sides of the chemical equation) as a unit.
4. Balance the remaining atoms, usually ending with the least complex substance and using fractional coefficients if
necessary. If a fractional coefficient has been used, multiply both sides of the equation by the denominator to obtain whole
numbers for the coefficients.
5. Count the numbers of atoms of each kind on both sides of the equation to be sure that the chemical equation is balanced.

 Example 3.8.1: Combustion of Heptane

Balance the chemical equation for the combustion of Heptane (C 7


H
16
).

C H (l) + O (g) → CO (g) + H O(g)


7 16 2 2 2

Solution
Solutions to Example 7.4.1
Steps Example

The most complex substance is the one with the largest number of
different atoms, which is C H . We will assume initially that the
1. Identify the most complex substance. 7 16

final balanced chemical equation contains 1 molecule or formula unit


of this substance.
a. Because one molecule of n-heptane contains 7 carbon atoms, we
need 7 CO2 molecules, each of which contains 1 carbon atom, on the
right side:

C H (l) + O (g) → 7CO (g) + H O(g)


7 16 2 – 2 2

7 carbon atoms on both reactant and product sides


2. Adjust the coefficients.
b. Because one molecule of n-heptane contains 16 hydrogen atoms,
we need 8 H2O molecules, each of which contains 2 hydrogen atoms,
on the right side:

C H (l) + O (g) → 7 CO (g) + 8H O(g)


7 16 2 2 – 2

16 hydrogen atoms on both reactant and product sides

3. Balance polyatomic ions as a unit. There are no polyatomic ions to be considered in this reaction.

The carbon and hydrogen atoms are now balanced, but we have 22
oxygen atoms on the right side and only 2 oxygen atoms on the left.
We can balance the oxygen atoms by adjusting the coefficient in front
4. Balance the remaining atoms. of the least complex substance, O2, on the reactant side:

C H (l) + 11O (g) → 7 CO (g) + 8 H O(g)


7 16 ––
– 2 2 2

22 oxygen atoms on both reactant and product sides

The equation is now balanced, and there are no fractional


coefficients: there are 7 carbon atoms, 16 hydrogen atoms, and 22
5. Check your work.
oxygen atoms on each side. Always check to be sure that a chemical
equation is balanced.

3.8.2 https://chem.libretexts.org/@go/page/423623
 Example 3.8.2: Combustion of Isooctane

Combustion of Isooctane (C 8
H
18
)

C H (l) + O (g) ⟶ CO (g) + H O(g)


8 18 2 2 2

Solution
The assumption that the final balanced chemical equation contains only one molecule or formula unit of the most complex
substance is not always valid, but it is a good place to start. The combustion of any hydrocarbon with oxygen produces carbon
dioxide and water.
Solutions to Example 7.4.2
Steps Example

The most complex substance is the one with the largest number of
different atoms, which is C H . We will assume initially that the
1. Identify the most complex substance. 8 18

final balanced chemical equation contains 1 molecule or formula unit


of this substance.
a. The first element that appears only once in the reactants is carbon:
8 carbon atoms in isooctane means that there must be 8 CO2
molecules in the products:

C H (l) + O (g) ⟶ 8CO (g) + H O(g)


8 18 2 – 2 2

8 carbon atoms on both reactant and product sides


2. Adjust the coefficients.
b. 18 hydrogen atoms in isooctane means that there must be 9 H2O
molecules in the products:

C H (l) + O (g) ⟶ 8 CO (g) + 9H O(g)


8 18 2 2 – 2

18 hydrogen atoms on both reactant and product sides

3. Balance polyatomic ions as a unit. There are no polyatomic ions to be considered in this reaction.

The carbon and hydrogen atoms are now balanced, but we have 25
oxygen atoms on the right side and only 2 oxygen atoms on the left.
We can balance the least complex substance, O2, but because there
are 2 oxygen atoms per O2 molecule, we must use a fractional
25
coefficient ( ) to balance the oxygen atoms:
2

25
C H (l) + O (g) → 8 CO (g) + 9 H O(g)
8 18 2 2 2
2
–––
4. Balance the remaining atoms.
25 oxygen atoms on both reactant and product sides
The equation is now balanced, but we usually write equations with
whole number coefficients. We can eliminate the fractional
coefficient by multiplying all coefficients on both sides of the
chemical equation by 2:

2C H (l) + 25O (g) ⟶ 16CO (g) + 18H O(g)


– 8 18 ––
– 2 ––
– 2 ––
– 2

The balanced chemical equation has 16 carbon atoms, 36 hydrogen


atoms, and 50 oxygen atoms on each side.
Balancing equations requires some practice on your part as well as
5. Check your work. some common sense. If you find yourself using very large
coefficients or if you have spent several minutes without success, go
back and make sure that you have written the formulas of the
reactants and products correctly.

3.8.3 https://chem.libretexts.org/@go/page/423623
 Example 3.8.3: Precipitation of Lead (II) Chloride
Aqueous solutions of lead (II) nitrate and sodium chloride are mixed. The products of the reaction are an aqueous solution of
sodium nitrate and a solid precipitate of lead (II) chloride. Write the balanced chemical equation for this reaction.

Solution
Solutions to Example 7.4.3
Steps Example

The most complex substance is lead (II) chloride.


1. Identify the most complex substance.
Pb(NO ) (aq) + NaCl(aq) → NaNO (aq) + PbCl (s)
3 2 3 2

There are twice as many chloride ions in the reactants as there are in
the products. Place a 2 in front of the NaCl in order to balance the
chloride ions.

2. Adjust the coefficients. Pb(NO ) (aq) + 2NaCl(aq) → NaNO (aq) + PbCl (s)
3 2 – 3 2

1 Pb atom on both reactant and product sides


2 Na atoms on reactant side, 1 Na atom on product side
2 Cl atoms on both reactant and product sides

The nitrate ions are still unbalanced. Place a 2 in front of the NaNO3.
The result is:

Pb(NO ) (aq) + 2 NaCl(aq) → 2NaNO (aq) + PbCl (s)


3 2 – 3 2

3. Balance polyatomic ions as a unit.


1 Pb atom on both reactant and product sides
2 Na atoms on both reactant and product sides
2 Cl atoms on both reactant and product sides
2 NO3- atoms on both reactant and product sides

There is no need to balance the remaining atoms because they are


4. Balance the remaining atoms.
already balanced.

Pb(NO ) (aq) + 2 NaCl(aq) → 2 NaNO (aq) + PbCl (s)


3 2 3 2

1 Pb atom on both reactant and product sides


5. Check your work.
2 Na atoms on both reactant and product sides
2 Cl atoms on both reactant and product sides
2 NO3- atoms on both reactant and product sides

 Exercise 3.8.1

Is each chemical equation balanced?


a. 2 Hg(ℓ) O (g) → Hg O (s)
+

2 2 2

b. C H (g) + 2 O (g) → 2 CO (g) + 2 H O(g)


2 4 2 2 2

c. Mg(NO ) (s) + 2 Li(s) → Mg(s) 2 LiNO (s)


3 2
+

Answer a
yes

Answer b
no

Answer c
yes

3.8.4 https://chem.libretexts.org/@go/page/423623
 Exercise 3.8.2

Balance the following chemical equations.


a. N (g) + O (g) → NO (g)
2 2 2

b. Pb(NO ) (aq) + FeCl (aq) → Fe(NO )


3 2 3 3 3
(aq) + PbCl (s)
2

c. C H (l) + O (g) → CO (g) + H O(g)


6 14 2 2 2

Answer a
N2 (g) + 2O2 (g) → 2NO2 (g)

Answer b
3Pb(NO3)2(aq) + 2FeCl3(aq) → 2Fe(NO3)3(aq) + 3PbCl2(s)

Answer c
2C6H14(l) + 19O2(g)→ 12CO2(g) + 14H2O(g)

Summary
To be useful, chemical equations must always be balanced. Balanced chemical equations have the same number and type of
each atom on both sides of the equation.
The coefficients in a balanced equation must be the simplest whole number ratio. Mass is always conserved in chemical
reactions.

Vocabulary
Chemical reaction - The process in which one or more substances are changed into one or more new substances.
Reactants - The starting materials in a reaction.
Products - Materials present at the end of a reaction.
Balanced chemical equation - A chemical equation in which the number of each type of atom is equal on the two sides of the
equation.
Subscripts - Part of the chemical formulas of the reactants and products that indicate the number of atoms of the preceding
element.
Coefficient - A small whole number that appears in front of a formula in a balanced chemical equation.

3.8: How to Write Balanced Chemical Equations is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-
Agnew & Henry Agnew.
7.4: How to Write Balanced Chemical Equations by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

3.8.5 https://chem.libretexts.org/@go/page/423623
3.9: Making Pancakes- Relationships Between Ingredients
3.9: Making Pancakes- Relationships Between Ingredients is shared under a not declared license and was authored, remixed, and/or curated by
LibreTexts.

3.9.1 https://chem.libretexts.org/@go/page/423624
3.10: MAKING MOLECULES- MOLE-TO-MOLE CONVERSIONS
 /;LEARNING OBJECTIVES
Use a balanced chemical equation to determine molar relationships between substances.

Previously, you learned to balance chemical equations by comparing the numbers of each type of atom in the reactants and products. The
coefficients in front of the chemical formulas represent the numbers of molecules or formula units (depending on the type of substance). As
follows, we will extend the meaning of the coefficients in a chemical equation.
Consider the simple chemical equation:
2H2 + O2 → 2H2 O

The convention for writing balanced chemical equations is to use the lowest whole-number ratio for the coefficients. However, the equation
is balanced as long as the coefficients are in a 2:1:2 ratio. For example, this equation is also balanced if we write it as

4H2 + 2O2 → 4H2 O

The ratio of the coefficients is 4:2:4, which reduces to 2:1:2. The equation is also balanced if we were to write it as

22H2 + 11O2 → 22H2 O

because 22:11:22 also reduces to 2:1:2.


Suppose we want to use larger numbers. Consider the following coefficients:
23 23 23
12.044 × 10 H + 6.022 × 10 O → 12.044 × 10 H O
2 2 2

These coefficients also have the ratio 2:1:2 (check it and see), so this equation is balanced. But 6.022 × 1023 is 1 mol, while 12.044 × 1023 is
2 mol (and the number is written that way to make this more obvious), so we can simplify this version of the equation by writing it as

2 mol H + 1 mol O → 2 mol H O


2 2 2

We can leave out the word mol and not write the 1 coefficient (as is our habit), so the final form of the equation, still balanced, is
2H +O → 2H O
2 2 2

Now we interpret the coefficients as referring to molar amounts, not individual molecules. The lesson? Balanced chemical equations are
balanced not only at the molecular level, but also in terms of molar amounts of reactants and products. Thus, we can read this reaction as
“two moles of hydrogen react with one mole of oxygen to produce two moles of water.”
By the same token, the ratios we constructed to describe a molecular reaction can also be constructed in terms of moles rather than
molecules. For the reaction in which hydrogen and oxygen combine to make water, for example, we can construct the following ratios:
2 mol H2 1 mol O2
or
1 mol O2 2 mol H2

2 mol H2 O 1 mol O2
or
1 mol O2 2 mol H2 O

2 mol H2 2 mol H2 O
or
2 mol H2 O 2 mol H2

We can use these ratios to determine what amount of a substance, in moles, will react with or produce a given number of moles of a
different substance. The study of the numerical relationships between the reactants and the products in balanced chemical reactions is called
stoichiometry.

 EXAMPLE 3.10.1

How many moles of oxygen react with hydrogen to produce 27.6 mol of H 2
O ?

Solution

3.10.1 https://chem.libretexts.org/@go/page/423625
Solutions to Example 8.3.1
Steps for Problem Solving How many moles of oxygen react with hydrogen to produce 27.6 mol of H 2
O ?
Unbalanced: H2 + O2 → H2O
Find a balanced equation that describes the reaction.
Balanced: 2H2 + O2 → 2H2O
Identify the "given" information and what the problem is
Given: moles H2O Find: moles oxygen
asking you to "find."
List other known quantities. 1 mol O2 = 2 mol H2O

Prepare a concept map and use the proper conversion


factor.

1 mol O2
27.6 mol H2 O × = 13.8 mol O2
Cancel units and calculate. 2 mol H2 O

To produce 27.6 mol of H2O, 13.8 mol of O2 react.


Since each mole of oxygen produces twice as many moles of water, it makes sense that the produced
Think about your result.
amount is greater than the reactant amount

 EXAMPLE 3.10.2

How many moles of ammonia are produced if 4.20 moles of hydrogen are reacted with an excess of nitrogen?

Solution
Solutions to Example 8.3.2
How many moles of ammonia are produced if 4.20 moles of hydrogen are reacted with an
Steps for Problem Solving
excess of nitrogen?
Unbalanced: N2 + H2 → NH3
Find a balanced equation that describes the reaction.
Balanced: N2 + 3H2 → 2NH3
Identify the "given" information and what the problem is asking Given: H = 4.20 mol
2

you to "find." Find: mol of NH 3

List other known quantities. 3 mol H2 = 2 mol NH3

Prepare a concept map and use the proper conversion factor.

2 mol N H3
4.20 mol H2 × = 2.80 mol N H3
Cancel units and calculate. 3 mol H2

The reaction of 4.20 mol of hydrogen with excess nitrogen produces 2.80 mol of ammonia.
Think about your result. The result corresponds to the 3:2 ratio of hydrogen to ammonia from the balanced equation.

 EXERCISE 3.10.3

a. Given the following balanced chemical equation:


C H +8O → 5 CO +6H O
5 12 2 2 2

, How many moles of H O can be formed if 0.0652 mol of C H were to react?


2 5 12

b. Balance the following unbalanced equation and determine how many moles of H 2
O are produced when 1.65 mol of NH3 react:

NH +O → N +H O
3 2 2 2

Answer a

3.10.2 https://chem.libretexts.org/@go/page/423625
0.391 mol H2O
Answer b
4NH3 + 3O2 → 2N2 + 6H2O; 2.48 mol H2O

SUMMARY
The balanced chemical reaction can be used to determine molar relationships between substances.

3.10: Making Molecules- Mole-to-Mole Conversions is shared under a CC BY-NC-SA 3.0 license and was authored, remixed, and/or curated by Marisa
Alviar-Agnew & Henry Agnew.
8.3: Making Molecules- Mole-to-Mole Conversions by Henry Agnew, Marisa Alviar-Agnew is licensed CC BY-NC-SA 3.0.

3.10.3 https://chem.libretexts.org/@go/page/423625
3.11: MAKING MOLECULES- MASS-TO-MASS CONVERSIONS
 LEARNING OBJECTIVES
Convert from mass or moles of one substance to mass or moles of another substance in a chemical reaction.

MOLE TO MASS CONVERSIONS


We have established that a balanced chemical equation is balanced in terms of moles, as well as atoms or molecules. We have used balanced
equations to set up ratios, in terms of moles of materials, that we can use as conversion factors to answer stoichiometric questions—such as
how many moles of substance A react with so many moles of reactant B. We can extend this technique even further. Recall that we can
relate a molar amount to a mass amount using molar mass. We can use that relation to answer stoichiometry questions in terms of the
masses of a particular substance, in addition to moles. We do this using the following sequence:

Collectively, these conversions are called mole-mass calculations. Flowchart of mole mass calculations: To convert from moles of substance
A to moles of substance B, use the mole ratio conversion factor, and to convert from moles to grams of substance B, use molar mass
conversion factor
As an example, consider the balanced chemical equation
Fe O + 3 SO → Fe (SO ) (3.11.1)
2 3 3 2 4 3

If we have 3.59 mol of Fe O , how many grams of SO can react with it? Using the mole-mass calculation sequence, we can determine the
2 3 3

required mass of SO in two steps. First, we construct the appropriate molar ratio, determined from the balanced chemical equation, to
3

calculate the number of moles of SO needed. Then, using the molar mass of SO as a conversion factor, we determine the mass that this
3 3

number of moles of SO has. 3

As usual, we start with the quantity we were given:

3 mol SO3
3.59 mol Fe2 O3 ×( ) = 10.77 mol SO3 (3.11.2)
1 mol Fe2 O3

The mol Fe O units cancel, leaving mol SO unit. Now, we take this answer and convert it to grams of SO , using the molar mass of SO
2 3 3 3 3

as the conversion factor:

80.06 g SO3
10.77 mol SO3 ×( ) = 862 g SO3 (3.11.3)
1 mol SO3

Our final answer is expressed to three significant figures. Thus, in a two-step process, we find that 862 g of SO will react with 3.59 mol of
3

Fe O . Many problems of this type can be answered in this manner.


2 3

The same two-step problem can also be worked out in a single line, rather than as two separate steps, as follows:

3.11.1 https://chem.libretexts.org/@go/page/423626
3 mol S O3 80.06 g S O3
3.59 mol F e2 O3 × ( )× ( ) = 862 g S O3
1 mol F e2 O3 1 mol S O3
 
converts to moles of SO 3 converts to grams of SO 3

We get exactly the same answer when combining all math steps together.

 EXAMPLE 3.11.1: GENERATION OF ALUMINUM OXIDE

How many moles of HCl will be produced when 249 g of AlCl are reacted according to this chemical equation?
3

2 AlCl 3 + 3 H2 O(ℓ) → Al 2 O3 + 6 HCl(g)

Solution
Solutions to Example 8.5.1
Steps for Problem Solving Example 3.11.1
Identify the "given" information and what the Given: 249 g AlCl3
problem is asking you to "find." Find: moles HCl
1 mol AlCl3 = 133.33 g AlCl3
List other known quantities.
6 mol of HCl to 2 mol AlCl3

Prepare a concept map and use the proper


conversion factor.

Flowchart of needed conversion factors: 1 mole AlCl3 to 133.33 grams AlCl3, and 6 moles HCl to 2 moles AlCl3

1 mol AlCl3 6 mol HCl


Cancel units and calculate. 249 g AlCl3 × × = 5.60 mol HCl
133.33 g AlCl3 2 mol AlCl3

Since 249 g of AlCl3 is less than 266.66 g, the mass for 2 moles of AlCl3 and the relationship is 6 mol of HCl to 2
Think about your result.
mol AlCl3 , the answer should be less than 6 moles of HCl.

 EXERCISE 3.11.1: GENERATION OF ALUMINUM OXIDE

How many moles of Al 2


O
3
will be produced when 23.9 g of H 2
O are reacted according to this chemical equation?

2 AlCl + 3 H O(ℓ) → Al O + 6 HCl(g)


3 2 2 3

Answer
0.442 mol Al2O3

MASS TO MASS CONVERSIONS


It is a small step from mole-mass calculations to mass-mass calculations. If we start with a known mass of one substance in a chemical
reaction (instead of a known number of moles), we can calculate the corresponding masses of other substances in the reaction. The first step
in this case is to convert the known mass into moles, using the substance’s molar mass as the conversion factor. Then—and only then—we
use the balanced chemical equation to construct a conversion factor to convert that quantity to moles of another substance, which in turn can
be converted to a corresponding mass. Sequentially, the process is as follows:

3.11.2 https://chem.libretexts.org/@go/page/423626
This three-part process can be carried out in three discrete steps or combined into a single calculation that contains three conversion factors.
The following example illustrates both techniques. Flowchart of mole mass calculations: To convert from grams to moles of substance A,
use molar mass conversion factor; To convert from moles of substance A to moles of substance B, use the mole ratio conversion factor, and
to convert from moles to grams of substance B, use molar mass conversion factor

 EXAMPLE 3.11.2: DECOMPOSITION OF AMMONIUM NITRATE

Ammonium nitrate decomposes to dinitrogen monoxide and water according to the following equation.

NH NO (s) → N O (g) + 2H O (l)


4 3 2 2

In a certain experiment, 45.7 g of ammonium nitrate is decomposed. Find the mass of each of the products formed.

3.11.3 https://chem.libretexts.org/@go/page/423626
Solutions to Example 8.5.2
Steps for Problem Solving Example 3.11.2
Given: 45.7 g NH 4
NO
3

Identify the "given" information and what the Find:


problem is asking you to "find." Mass N O =? g
2

Mass H O =? g
2

1 mol NH NO = 80.06 g
4 3

1 mol N O = 44.02 g
List other known quantities. 2

1 mol H O = 18.02 g
2

1 mol NH4NO3 to 1 mol N2O to 2 mol H2O

Flowchart of conversion factors: 1 mole NH4NO3 to 80.06 grams NH4NO3, 1 mole N2O to 1 mole NH4NO3, 44.02
Prepare two concept maps and use the proper grams N2O to 1 mole N2O
conversion factor.

Flowchart of conversion factors: 1 mole NH4NO3 to 80.06 grams NH4NO3, 2 moles H2O to 1 mole NH4NO3,
18.02 grams H2O to 1 mole H2O
1 mol NH NO 1 mol N O 44.02 g N O
4 3 2 2
45.7 g NH NO × × × = 25.1 g N O
4 3 2
80.06 g NH NO 1 mol NH NO 1 mol N O
Cancel units and calculate. 1 mol NH NO
4 3 4

2 mol H O
3 2

18.02 g H O
4 3 2 2
45.7 g NH NO × × × = 20.6 g H O
4 3 2
80.06 g NH NO 1 mol NH NO 1 mol H O
4 3 4 3 2

The total mass of the two products is equal to the mass of ammonium nitrate which decomposed, demonstrating the
Think about your result.
law of conservation of mass. Each answer has three significant figures.

 EXERCISE 3.11.2: CARBON TETRACHLORIDE

Methane can react with elemental chlorine to make carbon tetrachloride (CCl ). The balanced chemical equation is as follows:
4

CH (g) + 4 Cl (g) → CCl (l) + 4 HCl(l)


4 2 2

How many grams of HCl are produced by the reaction of 100.0 g of CH ? 4

Answer
908.7g HCl

SUMMARY
Calculations involving conversions between moles of a substance and the mass of that substance can be done using conversion factors.
A balanced chemical reaction can be used to determine molar and mass relationships between substances.

3.11: Making Molecules- Mass-to-Mass Conversions is shared under a CC BY-NC-SA 3.0 license and was authored, remixed, and/or curated by Marisa
Alviar-Agnew & Henry Agnew.
8.4: Making Molecules- Mass-to-Mass Conversions by Henry Agnew, Marisa Alviar-Agnew is licensed CC BY-NC-SA 3.0.

3.11.4 https://chem.libretexts.org/@go/page/423626
3.12: Stoichiometry
 Learning Objectives
Explain the meaning of the term "stoichiometry".
Determine the relative amounts of each substance in chemical equations.

You have learned that chemical equations provide us with information about the types of particles that react to form products.
Chemical equations also provide us with the relative number of particles and moles that react to form products. In this section you
will explore the quantitative relationships that exist between the quantities of reactants and products in a balanced equation. This is
known as stoichiometry.
Stoichiometry, by definition, is the calculation of the quantities of reactants or products in a chemical reaction using the
relationships found in the balanced chemical equation. The word stoichiometry is actually Greek from two words: στ οικηιον,
which means "element", and \(\mu \epsilon \tau \rho \omicron \nu), which means "measure".

Interpreting Chemical Equations


The mole, as you remember, is a quantitative measure that is equivalent to Avogadro's number of particles. So how does this relate
to the chemical equation? Look at the chemical equation below.

2 CuSO + 4KI → 2CuI + 2 K SO +I


4 2 4 2

The coefficients used, as we have learned, tell us the relative amounts of each substance in the equation. So for every 2 units of
copper (II) sulfate (CuSO ) we have, we need to have 4 units of potassium iodide (KI ). For every two dozen copper (II) sulfates,
4

we need 4 dozen potassium iodides. Because the unit "mole" is also a counting unit, we can interpret this equation in terms of
moles, as well: For every two moles of copper (II) sulfate, we need 4 moles potassium iodide.
The production of ammonia (NH 3
) from nitrogen and hydrogen gases is an important industrial reaction called the Haber process,
after German chemist Fritz Haber.

N (g) + 3 H (g) → 2 NH (g)


2 2 3

The balanced equation can be analyzed in several ways, as shown in the figure below.

Figure 3.12.1 : This representation of the production of ammonia from nitrogen and hydrogen show several ways to interpret the
quantitative information of a chemical reaction.
We see that 1 molecule of nitrogen reacts with 3 molecules of hydrogen to form 2 molecules of ammonia. This is the smallest
possible relative amount of the reactants and products. To consider larger relative amounts, each coefficient can be multiplied by
the same number. For example, 10 molecules of nitrogen would react with 30 molecules of hydrogen to produce 20 molecules of
ammonia.
The most useful quantity for counting particles is the mole. So if each coefficient is multiplied by a mole, the balanced chemical
equation tells us that 1 mole of nitrogen reacts with 3 moles of hydrogen to produce 2 moles of ammonia. This is the conventional
way to interpret any balanced chemical equation.
Finally, if each mole quantity is converted to grams by using the molar mass, we can see that the law of conservation of mass is
followed. 1 mol of nitrogen has a mass of 28.02 g, while 3 mol of hydrogen has a mass of 6.06 g, and 2 mol of ammonia has a
mass of 34.08 g.

3.12.1 https://chem.libretexts.org/@go/page/423627
28.02 g N + 6.06 g H → 34.08 g NH
2 2 3

Mass and the number of atoms must be conserved in any chemical reaction. The number of molecules is not necessarily conserved.

 Example 3.12.1

The equation for the combustion of ethane (C 2


H
6
) is

2C H + 7O → 4 CO + 6H O
2 6 2 2 2

a. Indicate the number of formula units or molecules in the balanced equation.


b. Indicate the number of moles present in the balanced equation.

Solution
a. Two molecules of C H plus seven molecules of O yields four molecules of CO plus six molecules of H
2 6 2 2 2
O .
b. Two moles of C H plus seven moles of O yields four moles of CO plus six moles of H O .
2 6 2 2 2

 Exercise 3.12.1

For the following equation below, indicate the number of formula units or molecules, and the number of moles present in the
balanced equation.

KBrO + 6KI + 6HBr → 7KBr + 3 H O


3 2

Answer
One molecules of KBrO plus six molecules of KI plus six molecules of HBr yields seven molecules of KBr plus three
3

molecules of I and three molecules of H O . One mole of KBrO plus six moles of KI plus six moles of HBr yields
2 2 3

seven moles of KBr plus three moles of I plus three moles of H O .


2 2

Summary
Stoichiometry is the calculation of the quantities of reactants or products in a chemical reaction using the relationships found in
a balanced chemical equation.
The coefficients in a balanced chemical equation represent the reacting ratios of the substances in the reaction.
The coefficients of a balanced equation can be used to determine the ratio of moles of all substances in the reaction.

Vocabulary
Stoichiometry - The calculation of quantitative relationships of the reactants and products in a balanced chemical equation.
Formula unit - The empirical formula of an ionic compound.
Mole ratio - The ratio of the moles of one reactant or product to the moles of another reactant or product according to the
coefficients in the balanced chemical equation.

3.12: Stoichiometry is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew & Henry Agnew.
8.5: Stoichiometry by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source: https://www.ck12.org/c/chemistry/.

3.12.2 https://chem.libretexts.org/@go/page/423627
3.13: LIMITING REACTANT AND THEORETICAL YIELD
 LEARNING OBJECTIVES
Identify the limiting reactant (limiting reagent) in a given chemical reaction.
Calculate how much product will be produced from the limiting reactant.
Calculate how much reactant(s) remains when the reaction is complete.

In all examples discussed thus far, the reactants were assumed to be present in stoichiometric quantities. Consequently, none of the reactants were left over at the end of the reaction. This is often desirable—as in
the case of a space shuttle—where excess oxygen or hydrogen is not only extra freight to be hauled into orbit, but also an explosion hazard. More often, however, reactants are present in mole ratios that are not the
same as the ratio of the coefficients in the balanced chemical equation. As a result, one or more of them will not be used up completely, but will be left over when the reaction is completed. In this situation, the
amount of product that can be obtained is limited by the amount of only one of the reactants. The reactant that restricts the amount of product obtained is called the limiting reactant. The reactant that remains after
a reaction has gone to completion is in excess.
Consider a nonchemical example. Assume you have invited some friends for dinner and want to bake brownies for dessert. You find two boxes of brownie mix in your pantry and see that each package requires
two eggs. The balanced equation for brownie preparation is:
1 box mix + 2 eggs → 1 batch brownies (3.13.1)

If you have a dozen eggs, which ingredient will determine the number of batches of brownies that you can prepare? Because each box of brownie mix requires two eggs and you have two boxes, you need four
eggs. Twelve eggs is eight more eggs than you need. Although the ratio of eggs to boxes in is 2:1, the ratio in your possession is 6:1. Hence the eggs are the ingredient (reactant) present in excess, and the brownie
mix is the limiting reactant. Even if you had a refrigerator full of eggs, you could make only two batches of brownies.

Figure 3.13.1: The Concept of a Limiting Reactant in the Preparation of Brownies. For a chemist, the balanced chemical equation is the recipe that must be followed.
2 boxes of brownie mix and 12 eggs results in 2 batches of brownies and 8 eggs; in this case the 8 eggs are reactant present in excess

 PHET SIMULATION: REACTANTS, PRODUCTS AND LEFTOVERS

View this interactive simulation illustrating the concepts of limiting and excess reactants.

Reactants, Products and Leftovers

Molecules Game

Sandwiches

Consider this concept now with regard to a chemical process, the reaction of hydrogen with chlorine to yield hydrogen chloride:

H + Cl (g) → 2 HCl(g)
2 2

The balanced equation shows that hydrogen and chlorine react in a 1:1 stoichiometric ratio. If these reactants are provided in any other amounts, one of the reactants will nearly always be entirely consumed, thus
limiting the amount of product that may be generated. This substance is the limiting reactant, and the other substance is the excess reactant. Identifying the limiting and excess reactants for a given situation
requires computing the molar amounts of each reactant provided and comparing them to the stoichiometric amounts represented in the balanced chemical equation. For example, imagine combining 3 moles of H2

3.13.1 https://chem.libretexts.org/@go/page/423628
and 2 moles of Cl2. This represents a 3:2 (or 1.5:1) ratio of hydrogen to chlorine present for reaction, which is greater than the stoichiometric ratio of 1:1. Hydrogen, therefore, is present in excess, and chlorine is
the limiting reactant. Reaction of all the provided chlorine (2 mol) will consume 2 mol of the 3 mol of hydrogen provided, leaving 1 mol of hydrogen non-reacted.
An alternative approach to identifying the limiting reactant involves comparing the amount of product expected for the complete reaction of each reactant. Each reactant amount is used to separately calculate the
amount of product that would be formed per the reaction’s stoichiometry. The reactant yielding the lesser amount of product is the limiting reactant. For the example, in the previous paragraph, complete reaction
of the hydrogen would yield:
2 mol HCl
mol HCl produced = 3 mol H2 × = 6 mol HCl
1 mol H2

Complete reaction of the provided chlorine would produce:


2 mol HCl
mol HCl produced = 2 mol Cl2 × = 4 mol HCl
1 mol Cl2

The chlorine will be completely consumed once 4 moles of HCl have been produced. Since enough hydrogen was provided to yield 6 moles of HCl, there will be non-reacted hydrogen remaining once this reaction
is complete. Chlorine, therefore, is the limiting reactant and hydrogen is the excess reactant (Figure 3.13.2 ).

Figure 3.13.2: When H2 and Cl2 are combined in nonstoichiometric amounts, one of these reactants will limit the amount of HCl that can be produced. This illustration shows a reaction in which hydrogen is
present in excess and chlorine is the limiting reactant. The figure shows a space-filling molecular models reacting. There is a reaction arrow pointing to the right in the middle. To the left of the reaction arrow there
are three molecules each consisting of two green spheres bonded together. There are also five molecules each consisting of two smaller, white spheres bonded together. Above these molecules is the label, “Before
reaction,” and below these molecules is the label, “6 H subscript 2 and 4 C l subscript 2.” To the right of the reaction arrow, there are eight molecules each consisting of one green sphere bonded to a smaller white
sphere. There are also two molecules each consisting of two white spheres bonded together. Above these molecules is the label, “After reaction,” and below these molecules is the label, “8 H C l and 2 H subscript
2.”
A similar situation exists for many chemical reactions: you usually run out of one reactant before all of the other reactant has reacted. The reactant you run out of is called the limiting reactant; the other reactant or
reactants are considered to be in excess. A crucial skill in evaluating the conditions of a chemical process is to determine which reactant is the limiting reactant and which is in excess.

 HOW TO IDENTIFY THE LIMITING REACTANT (LIMITING REAGENT)

There are two ways to determine the limiting reactant. One method is to find and compare the mole ratio of the reactants used in the reaction (Approach 1). Another way is to calculate the grams of products
produced from the given quantities of reactants; the reactant that produces the smallest amount of product is the limiting reactant (Approach 2). This section will focus more on the second method.
Approach 1 (The "Reactant Mole Ratio Method"): Find the limiting reactant by looking at the number of moles of each reactant.
1. Determine the balanced chemical equation for the chemical reaction.
2. Convert all given information into moles (most likely, through the use of molar mass as a conversion factor).
3. Calculate the mole ratio from the given information. Compare the calculated ratio to the actual ratio.
4. Use the amount of limiting reactant to calculate the amount of product produced.
5. If necessary, calculate how much is left in excess of the non-limiting (excess) reactant.
Approach 2 (The "The Product Method"): Find the limiting reactant by calculating and comparing the amount of product that each reactant will produce.
1. Balance the chemical equation for the chemical reaction.
2. Convert the given information into moles.
3. Use stoichiometry for each individual reactant to find the mass of product produced.
4. The reactant that produces a lesser amount of product is the limiting reactant.
5. The reactant that produces a larger amount of product is the excess reactant.
6. To find the amount of remaining excess reactant, subtract the mass of excess reactant consumed from the total mass of excess reactant given.

The key to recognizing which reactant is the limiting reactant is based on a mole-mass or mass-mass calculation: whichever reactant gives the lesser amount of product is the limiting reactant. What we need to do
is determine an amount of one product (either moles or mass) assuming all of each reactant reacts. Whichever reactant gives the least amount of that particular product is the limiting reactant. It does not matter
which product we use, as long as we use the same one each time. It does not matter whether we determine the number of moles or grams of that product; however, we will see shortly that knowing the final mass of
product can be useful.

 EXAMPLE 3.13.1: IDENTIFYING THE LIMITING REACTANT

As an example, consider the balanced equation

4 C2 H3 Br3 + 11 O2 → 8 CO2 + 6 H2 O + 6 Br2

What is the limiting reactant if 76.4 grams of C 2


H3 Br3 reacted with 49.1 grams of O ? 2

Solution
Using Approach 1:

Step 1: Balance the chemical equation.


The equation is already balanced with the relationship
4 mol C 2
H Br
3 3
to 11 mol O to 6 mol H
2 2
O to 6 mol Br

Step 2: Convert all given information into moles.


1 mol C2 H3 Br3
76.4 g C2 H3 Br3 × = 0.286 mol C2 H3 Br3
266.72 g C2 H3 B3

1 mol O2
49.1 g O2 × = 1.53 mol O2
32.00 g O2

3.13.2 https://chem.libretexts.org/@go/page/423628
Step 3: Calculate the mole ratio from the given information. Compare the calculated ratio to the actual ratio.
Assuming that all of the oxygen is used up,
4 molC2 H3 Br3
1.53 molO2 × = 0.556 mol C2H3Br3 are required.
11 molO2

Because 0.556 moles of C2H3Br3 required > 0.286 moles of C2H3Br3 available, C2H3Br3 is the limiting reactant.
Using Approach 2:

Step 1: Balance the chemical equation.


The equation is already balanced with the relationship
4 mol C2H3Br3 to 11 mol O2 to 6 mol H2O to 6 mol Br2

Step 2 and Step 3: Convert mass to moles and stoichiometry.


1 mol C2 H3 Br3 8 mol CO2 44.01 g CO2
76.4 g C2 H3 Br3 × × × = 25.2 g CO2
266.72 g C2 H3 Br3 4 mol C2 H3 Br3 1 mol CO2

1 mol O2 8 mol CO2 44.01 g CO2


49.1 g O2 × × × = 49.1 g CO2
32.00 g O2 11 mol O2 1 mol CO2

Step 4: The reactant that produces a smaller amount of product is the limiting reactant.
Therefore, by either method, C 2
H3 Br3 is the limiting reactant.

 EXAMPLE 3.13.2: IDENTIFYING THE LIMITING REACTANT AND THE MASS OF EXCESS REACTANT

For example, in the reaction of magnesium metal and oxygen, calculate the mass of magnesium oxide that can be produced if 2.40 g M g reacts with 10.0 g O . Also determine the amount of excess reactant.
2

MgO is the only product in the reaction.

Solution
Following Approach 1:

Step 1: Balance the chemical equation.


2 Mg (s) + O2 (g) → 2 MgO (s)
The balanced equation provides the relationship of 2 mol Mg to 1 mol O2 to 2 mol MgO

Step 2 and Step 3: Convert mass to moles and stoichiometry.


1 mol Mg 2 mol MgO 40.31 g MgO
2.40 g Mg × × × = 3.98 g MgO
24.31 g Mg 2 mol Mg 1 mol MgO

1 mol O2 2 mol MgO 40.31 g MgO


10.0 g O2 × × × = 25.2 g MgO
32.00 g O2 1 mol O2 1 mol MgO

Step 4: The reactant that produces a smaller amount of product is the limiting reactant.
Mg produces less MgO than does O2 (3.98 g MgO vs. 25.2 g MgO), therefore Mg is the limiting reactant in this reaction.
Step 5: The reactant that produces a larger amount of product is the excess reactant.
O2 produces more amount of MgO than Mg (25.2g MgO vs. 3.98 MgO), therefore O2 is the excess reactant in this reaction.
Step 6: Find the amount of remaining excess reactant by subtracting the mass of the excess reactant consumed from the total mass of excess reactant given.
Mass of excess reactant calculated using the limiting reactant:

1 mol Mg 1 mol O2 32.00 g O2


2.40 g Mg × × × = 1.58 g O2
24.31 g Mg 2 mol Mg 1 mol O2

OR
Mass of excess reactant calculated using the mass of the product:

1 mol MgO 1 mol O2 32.0 g O2


3.98 g MgO × × × = 1.58 g O2
40.31 g MgO 2 mol MgO 1 mol O2

Mass of total excess reactant given – mass of excess reactant consumed in the reaction:
10.0g O2 - (available) 1.58g O2 (used) = 8.42g O2 (excess)

Therefore, O2 is in excess.

 EXAMPLE 3.13.3: LIMITING REACTANT

What is the limiting reactant if 78.0 grams of Na2O2 were reacted with 29.4 grams of H2O? The unbalanced chemical equation is

Na O (s) + H O(l) → NaOH(aq) + H O (l)


2 2 2 2 2

3.13.3 https://chem.libretexts.org/@go/page/423628
Solution
Solutions to Example 8.4.3
Steps for Problem Solving- The Product Method Example 3.13.1
Given: 78.0 grams of Na2O2
Identify the "given" information and what the problem is asking you to "find." 29.4 g H2O
Find: limiting reactant
1 mol Na2O2= 77.96 g/mol
List other known quantities. 1 mol H2O = 18.02 g/mol
Since the amount of product in grams is not required, only the molar mass of the reactants is needed.
Na2O2 (s) + 2H2O (l) → 2NaOH (aq) + H2O2 (l)
Balance the equation.
The balanced equation provides the relationship of 1 mol Na2O2 to 2 mol H2O 2mol NaOH to 1 mol H2O2

Prepare a concept map and use the proper conversion factor.

Because the question only asks for the limiting reactant, we can perform two mass-mole calculations and determine which amount is less.

1 mol Na2 O2 2 mol NaOH 40 g NaOH


78.0 g N a2 O2 × × × = 2.00 mol NaOH
77.96 g Na2 O2 1 mol Na2 O2 1 mol NaOH

Cancel units and calculate. 1 mol H2 O 2 mol NaOH 40 g NaOH


29.4 g H2 O × × × = 1.63 mol NaOH
18.02 g H2 O 2 mol Na2 O2 1 mol NaOH

Therefore, H2O is the limiting reactant.

Think about your result.

 EXAMPLE 3.13.4: LIMITING REACTANT AND MASS OF EXCESS REACTANT

A 5.00 g quantity of Rb is combined with 3.44 g of MgCl according to this chemical reaction:
2

2Rb(s) + M gC l2 (s) → M g(s) + 2RbCl(s)

What mass of Mg is formed, and what mass of remaining reactant is left over?

Solution

3.13.4 https://chem.libretexts.org/@go/page/423628
Solutions to Example 8.4.10
Steps for Problem Solving- The Product
Example 3.13.2
Method
A 5.00 g quantity of Rb is combined with 3.44 g of MgCl2 according to this chemical reaction:

Steps for Problem Solving 2Rb(s) + M gC l2 (s) → M g(s) + 2RbCl(s)

What mass of Mg is formed, and what mass of remaining reactant is left over?
Identify the "given" information and what the Given: 5.00g Rb, 2.44g MgCl2
problem is asking you to "find." Find: mass of Mg formed, mass of remaining reactant
molar mass: Rb = 85.47 g/mol
List other known quantities. molar mass: MgCl2 = 95.21 g/mol
molar mass: Mg = 24.31 g/mol

Find mass Mg formed based on mass of Rb

Find mass of Mg formed based on mass of MgCl2

Prepare concept maps and use the proper


conversion factor.

Conversion factors: 1 mole MgCl2 to 95.21 grams MgCl2, 1 mole Mg to 1 mole MgCl2, 24.31 grams Mg to 1 mole Mg
Use limiting reactant to determine amount of excess reactant consumed

Conversion factors: 1 mole Rb to 85.47 grams Rb, 1 mole MgCl2 to 2 moles Rb, 95.21 grams MgCl2 to 1 mole MgCl2

Because the question asks what mass of magnesium is formed, we can perform two mass-mass calculations and determine which amount is less.

1 mol Rb 1 mol M g
24.31 g M g
5.00 g Rb × × × = 0.711 g M g
85.47 g Rb 2 mol Rb 1 mol M g

1 mol M gCl2 1 mol M g


24.31 g M g
3.44 g M gCl2 × × × = 0.878 g M g
95.21 g M gCl2 1 mol M gCl2 1 mol M g

Cancel units and calculate.


The 0.711 g of Mg is the lesser quantity, so the associated reactant—5.00 g of Rb—is the limiting reactant. To determine how much of the other reactant is left, we have to do one more mass-mass calculation
to determine what mass of MgCl2 reacted with the 5.00 g of Rb, and then subtract the amount reacted from the original amount.

1 mol Rb 1 mol M gCl2


95.21 g M gCl2
5.00 g Rb × × × = 2.78 g M gC l2 reacted
85.47 g Rb 2 mol Rb 1 mol M gCl2

Because we started with 3.44 g of MgCl2, we have


3.44 g MgCl2 − 2.78 g MgCl2 reacted = 0.66 g MgCl2 left
Think about your result. It usually is not possible to determine the limiting reactant using just the initial masses, as the reagents have different molar masses and coefficients.

 EXERCISE 3.13.1

Given the initial amounts listed, what is the limiting reactant, and what is the mass of the leftover reactant?

22.7 g + 17.9 g → M gS (s) + H2 O(l)


 
M gO(s) H2 S

Answer
H2S is the limiting reagent; 1.5 g of MgO are left over.

3.13: Limiting Reactant and Theoretical Yield is shared under a CC BY-NC-SA 3.0 license and was authored, remixed, and/or curated by Sarick Shah, Marisa Alviar-Agnew, Henry Agnew, & Henry Agnew.
8.6: Limiting Reactant and Theoretical Yield by Henry Agnew, Marisa Alviar-Agnew, Sarick Shah is licensed CC BY-NC-SA 3.0.

3.13.5 https://chem.libretexts.org/@go/page/423628
3.13.6 https://chem.libretexts.org/@go/page/423628
3.13.7 https://chem.libretexts.org/@go/page/423628
3.14: LIMITING REACTANT, THEORETICAL YIELD, AND PERCENT YIELD FROM
INITIAL MASSES OF REACTANTS
 LEARNING OBJECTIVES
Calculate percentage or actual yields from known amounts of reactants.

The world of pharmaceutical production is an expensive one. Many drugs have several steps in their synthesis and use costly chemicals. A
great deal of research takes place to develop better ways to make drugs faster and more efficiently. Studying how much of a compound is
produced in any given reaction is an important part of cost control.

PERCENT YIELD
Chemical reactions in the real world don't always go exactly as planned on paper. In the course of an experiment, many things will
contribute to the formation of less product than predicted. Besides spills and other experimental errors, there are usually losses due to an
incomplete reaction, undesirable side reactions, etc. Chemists need a measurement that indicates how successful a reaction has been. This
measurement is called the percent yield.
To compute the percent yield, it is first necessary to determine how much of the product should be formed based on stoichiometry. This is
called the theoretical yield, the maximum amount of product that can be formed from the given amounts of reactants. The actual yield is
the amount of product that is actually formed when the reaction is carried out in the laboratory. The percent yield is the ratio of the actual
yield to the theoretical yield, expressed as a percentage.
Actual Yield
Percent Yield = × 100%
Theoretical Yield

Percent yield is very important in the manufacture of products. Much time and money is spent improving the percent yield for chemical
production. When complex chemicals are synthesized by many different reactions, one step with a low percent yield can quickly cause a
large waste of reactants and unnecessary expense.
Typically, percent yields are understandably less than 100% because of the reasons indicated earlier. However, percent yields greater than
100% are possible if the measured product of the reaction contains impurities that cause its mass to be greater than it actually would be if

the product was pure. When a chemist synthesizes a desired chemical, he or she is always careful to purify the products of the reaction.
Example 3.14.1 illustrates the steps for determining percent yield.

 EXAMPLE 3.14.1: DECOMPOSITION OF POTASSIUM CHLORATE


Potassium chlorate decomposes upon slight heating in the presence of a catalyst, according to the reaction below:

2KClO (s) → 2KCl (s) + 3O (g)


3 2

In a certain experiment, 40.0 g KClO is heated until it completely decomposes. The experiment is performed and the oxygen gas is
3

collected and its mass is found to be 14.9 g .


a. What is the theoretical yield of oxygen gas?
b. What is the percent yield for the reaction?

Solution
a. Calculation of theoretical yield
First, we will calculate the theoretical yield based on the stoichiometry.

Step 1: Identify the "given" information and what the problem is asking you to "find".
Given: Mass of KClO 3
= 40.0 g

Mass of O2 collected = 14.9g


Find: Theoretical yield, g O2

3.14.1 https://chem.libretexts.org/@go/page/423629
Step 2: List other known quantities and plan the problem.
1 mol KClO3 = 122.55 g/mol
1 mol O2 = 32.00 g/mol

Step 3: Apply stoichiometry to convert from the mass of a reactant to the mass of a product:

Step 4: Solve.
1 mol KClO 3 mol O
3 2 32.00 g O
2
40.0 g KClO × × × = 15.7 g O
3 2
122.55 g KClO 2 mol KClO 1 mol O
3 3 2

The theoretical yield of O is 15.7 g , 15.67 g unrounded.


2

Step 5: Think about your result.


The mass of oxygen gas must be less than the 40.0 g of potassium chlorate that was decomposed.
b. Calculation of percent yield
Now we will use the actual yield and the theoretical yield to calculate the percent yield.

Step 1: Identify the "given" information and what the problem is asking you to "find".
Given: Theoretical yield =15.67 g, use the un-rounded number for the calculation.
Actual yield = 14.9g
Find: Percent yield, % Yield

Step 2: List other known quantities and plan the problem.

No other quantities needed.

Step 3: Use the percent yield equation below.


Actual Yield
Percent Yield = × 100%
Theoretical Yield

Step 4: Solve.
14.9 g
Percent Yield = × 100% = 94.9%
15.67 g

3.14.2 https://chem.libretexts.org/@go/page/423629
Step 5: Think about your result.
Since the actual yield is slightly less than the theoretical yield, the percent yield is just under 100% .

 EXAMPLE 3.14.2: OXIDATION OF ZINC

Upon reaction of 1.274 g of copper sulfate with excess zinc metal, 0.392 g copper metal was obtained according to the equation:
CuSO (aq) + Zn(s) → Cu(s) + ZnSO (aq)
4 4

What is the percent yield?

Solution
Solutions to Example 8.6.2
Steps for Problem Solving-The Product
Example 3.14.1
Method
Given: 1.274 g CuSO4
Identify the "given" information and what the Actual yield = 0.392 g Cu
problem is asking you to "find."
Find: Percent yield
1 mol CuSO4= 159.62 g/mol
List other known quantities. 1 mol Cu = 63.55 g/mol
Since the amount of product in grams is not required, only the molar mass of the reactants is needed.
The chemical equation is already balanced.
Balance the equation.
The balanced equation provides the relationship of 1 mol CuSO4 to 1 mol Zn to 1 mol Cu to 1 mol ZnSO4.

Prepare a concept map and use the proper


conversion factor.

The provided information identifies copper sulfate as the limiting reactant, and so the theoretical yield (g Cu) is found
by performing mass-mass calculation based on the initial amount of CuSO4.

1 mol CuSO4 1 mol Cu 63.55 g Cu


1.274 g CuS O4 × × × = 0.5072 g Cu
159.62 g CuSO4 1 mol CuSO4 1 mol Cu

Using this theoretical yield and the provided value for actual yield, the percent yield is calculated to be:
Cancel units and calculate. actual yield
percent yield = ( ) × 100
theoretical yield

0.392 g Cu
= ( ) × 100
0.5072 g Cu

= 77.3%

Think about your result. Since the actual yield is slightly less than the theoretical yield, the percent yield is just under 100%.

 EXERCISE 3.14.1

What is the percent yield of a reaction that produces 12.5 g of the Freon CF2Cl2 from 32.9 g of CCl4 and excess HF?

CCl4 + 2 HF → CF2 Cl2 + 2 HCl

Answer
48.3%

SUMMARY
Theoretical yield is calculated based on the stoichiometry of the chemical equation. The actual yield is experimentally determined. The
percent yield is determined by calculating the ratio of actual yield to theoretical yield.

3.14.3 https://chem.libretexts.org/@go/page/423629
3.14: Limiting Reactant, Theoretical Yield, and Percent Yield from Initial Masses of Reactants is shared under a CC BY-NC-SA 3.0 license and was
authored, remixed, and/or curated by Marisa Alviar-Agnew & Henry Agnew.
8.7: Limiting Reactant, Theoretical Yield, and Percent Yield from Initial Masses of Reactants by Henry Agnew, Marisa Alviar-Agnew is licensed
CC BY-NC-SA 3.0.

3.14.4 https://chem.libretexts.org/@go/page/423629
3.15: Enthalpy Change is a Measure of the Heat Evolved or Absorbed
When physical or chemical changes occur, they are generally accompanied by a transfer of energy. The law of conservation of
energy states that in any physical or chemical process, energy is neither created nor destroyed. In other words, the entire energy in
the universe is conserved. In order to better understand the energy changes taking place during a reaction, we need to define two
parts of the universe, called the system and the surroundings. The system is the specific portion of matter in a given space that is
being studied during an experiment or an observation. The surroundings are everything in the universe that is not part of the
system. In practical terms for a laboratory chemist, the system is the particular chemicals being reacted, while the surroundings is
the immediate vicinity within the room. During most processes, energy is exchanged between the system and the surroundings. If
the system loses a certain amount of energy, that same amount of energy is gained by the surroundings. If the system gains a certain
amount of energy, that energy is supplied by the surroundings.
A chemical reaction or physical change is endothermic if heat is absorbed by the system from the surroundings. In the course of an
endothermic process, the system gains heat from the surroundings and so the temperature of the surroundings decreases. The
quantity of heat for a process is represented by the letter q. The sign of q for an endothermic process is positive because the system
is gaining heat. A chemical reaction or physical change is exothermic if heat is released by the system into the surroundings.
Because the surroundings are gaining heat from the system, the temperature of the surroundings increases. The sign of q for an
exothermic process is negative because the system is losing heat.

Figure 3.15.1 : (A) Endothermic reaction. (B) Exothermic reaction.

Enthalpy
Heat changes in chemical reactions are often measured in the laboratory under conditions in which the reacting system is open to
the atmosphere. In that case, the system is at a constant pressure. Enthalpy (H) is the heat content of a system at constant
pressure. Chemists routinely measure changes in enthalpy of chemical systems as reactants are converted into products. The heat
that is absorbed or released by a reaction at constant pressure is the same as the enthalpy change, and is given the symbol ΔH .
Unless otherwise specified, all reactions in this material are assumed to take place at constant pressure.
The change in enthalpy of a reaction is a measure of the differences in enthalpy of the reactants and products. The enthalpy of a
system is determined by the energies needed to break chemical bonds and the energies needed to form chemical bonds. Energy
needs to be put into the system in order to break chemical bonds, as they do not come apart spontaneously in most cases. Bond
formation to produce products will involve release of energy. The change in enthalpy shows the trade-offs made in these two
processes. Does it take more energy to break bonds than that needed to form bonds? If so, the reaction is endothermic and the
enthalpy change is positive. If more energy is produced in bond formation than that needed for bond breaking, the reaction is
exothermic and the enthalpy is negative.
Several factors influence the enthalpy of a system. Enthalpy is an extensive property, determined in part by the amount of material
we work with. The state of reactants and products (solid, liquid, or gas) influences the enthalpy value for a system. The direction of
the reaction affects the enthalpy value. A reaction that takes place in the opposite direction has the same numerical enthalpy value,
but the opposite sign.

Thermochemical Equation
When methane gas is combusted, heat is released, making the reaction exothermic. Specifically, the combustion of 1 mol of
methane releases 890.4 kilojoules of heat energy. This information can be shown as part of the balanced equation:

CH (g) + 2 O (g) → CO (g) + 2 H O (l) + 890.4 kJ


4 2 2 2

3.15.1 https://chem.libretexts.org/@go/page/423630
The equation tells us that 1 mol of methane combines with 2 mol of oxygen to produce 1 mol of carbon dioxide and 2 mol of
water. In the process, 890.4 kJ is released and so it is written as a product of the reaction. A thermochemical equation is a
chemical equation that includes the enthalpy change of the reaction. The process in the above thermochemical equation can be
shown visually in Figure 3.15.2.

Figure 3.15.2 : (A) As reactants are converted to products in an exothermic reaction, enthalpy is released into the surroundings. The
enthalpy change of the reaction is negative. (B) As reactants are converted to products in an endothermic reaction, enthalpy is
absorbed from the surroundings. The enthalpy change of the reaction is positive.
In the combustion of methane example, the enthalpy change is negative because heat is being released by the system. Therefore,
the overall enthalpy of the system decreases. The heat of reaction is the enthalpy change for a chemical reaction. In the case
above, the heat of reaction is −890.4 kJ. The thermochemical reaction can also be written in this way:

CH (g) + 2 O (g) → CO (g) + 2 H O (l) ΔH = −890.4 kJ


4 2 2 2

Heats of reaction are typically measured in kilojoules. It is important to include the physical states of the reactants and products in a
thermochemical equation as the value of the ΔH depends on those states.
Endothermic reactions absorb energy from the surroundings as the reaction occurs. When 1 mol of calcium carbonate decomposes
into 1 mol of calcium oxide and 1 mol of carbon dioxide, 177.8 kJ of heat is absorbed. The process is shown visually in Figure
3.15.2B. The thermochemical reaction is shown below.

CaCO (s) + 177.8 kJ → CaO (s) + CO (g)


3 2

Because the heat is absorbed by the system, the 177.8 kJ is written as a reactant. The heat of reaction is positive for an
endothermic reaction.

CaCO (s) → CaO (s) + CO (g) ΔH = 177.8 kJ


3 2

The way in which a reaction is written influences the value of the enthalpy change for the reaction. Many reactions are reversible,
meaning that the product(s) of the reaction are capable of combining and reforming the reactant(s). If a reaction is written in the
reverse direction, the sign of the ΔH changes. For example, we can write an equation for the reaction of calcium oxide with carbon
dioxide to form calcium carbonate.

CaO (s) + CO (g) → CaCO (s) + 177.8 kJ


2 3

The reaction is exothermic and thus the sign of the enthalpy change is negative.

CaO (s) + CO (g) → CaCO (s) ΔH = −177.8 kJ


2 3

Stoichiometric Calculations and Enthalpy Changes


Chemistry problems that involve enthalpy changes can be solved by techniques similar to stoichiometry problems. Refer again to
the combustion reaction of methane. Since the reaction of 1 mol of methane released 890.4 kJ, the reaction of 2 mol of methane
890, 4 kJ
would release 2 × 890.4 kJ = 1781 kJ. The reaction of 0.5 mol of methane would release = 445.2 kJ . As with other
2
stoichiometry problems, the moles of a reactant or product can be linked to mass or volume.

 Example 3.15.1

Sulfur dioxide gas reacts with oxygen to form sulfur trioxide in an exothermic reaction, according to the following
thermochemical equation.

2 SO (g) + O (g) → 2 SO (g) + 198 kJ


2 2 3

3.15.2 https://chem.libretexts.org/@go/page/423630
Calculate the enthalpy change that occurs when 58.0 g of sulfur dioxide is reacted with excess oxygen.

Solution

Step 1: List the known quantities and plan the problem.


Mass SO = 58.0 g
2

Molar mass SO = 64.07 g/mol


2

ΔH = −198 kJ for the reaction of 2 mol SO


2

Unknown
ΔH =? kJ

The calculation requires two steps. The mass of SO


2
is converted to moles. Then the moles of SO
2
is multiplied by the
−198 kJ
conversion factor of ( .
)
2 mol SO
2

Step 2: Solve.
1 mol SO −198 kJ
2
ΔH = 58.0 g SO × × = 89.6 kJ
2
64.07 g SO 2 mol SO
2 2

Step 3: Think about your result.


The mass of sulfur dioxide is slightly less than 1 mol. Since 198 kJ is released for every 2 mol of SO that reacts, the heat 2

released when about 1 mol reacts is one half of 198. The 89.6 kJ is slightly less than half of 198. The sign of ΔH is negative
because the reaction is exothermic.

3.15: Enthalpy Change is a Measure of the Heat Evolved or Absorbed is shared under a CK-12 license and was authored, remixed, and/or curated
by Marisa Alviar-Agnew & Henry Agnew.
8.8: Enthalpy Change is a Measure of the Heat Evolved or Absorbed by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original
source: https://www.ck12.org/c/chemistry/.

3.15.3 https://chem.libretexts.org/@go/page/423630
CHAPTER OVERVIEW

4: Module 4 - Gases and Liquids


4.1: Kinetic Molecular Theory- A Model for Gases
4.2: Pressure - The Result of Constant Molecular Collisions
4.3: Boyle’s Law - Pressure and Volume
4.4: Charles’s Law- Volume and Temperature
4.5: Gay-Lussac's Law- Temperature and Pressure
4.6: Avogadro’s Law- Volume and Moles
4.7: The Ideal Gas Law- Pressure, Volume, Temperature, and Moles
4.8: Mixtures of Gases - Why Deep-Sea Divers Breathe a Mixture of Helium and Oxygen
4.9: Interactions between Molecules
4.10: Properties of Liquids and Solids
4.11: Intermolecular Forces in Action- Surface Tension and Viscosity
4.12: Evaporation and Condensation
4.13: Melting, Freezing, and Sublimation
4.14: Intermolecular Forces- Dispersion, Dipole–Dipole, Hydrogen Bonding, and Ion-Dipole
4.15: Water - A Remarkable Molecule

4: Module 4 - Gases and Liquids is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.

1
4.1: Kinetic Molecular Theory- A Model for Gases
 Learning Objectives
State the major concepts behind the kinetic theory of gases.
Relate the general properties of gases to the kinetic theory.

Gases were among the first substances studied in terms of the modern scientific method, which was developed in the 1600s. It did
not take long to recognize that gases all shared certain physical behaviors, suggesting that all gases could be described by one all-
encompassing theory. Today, that theory is the kinetic theory of gases. It is based on the following statements:
1. Gases consist of tiny particles of matter that are in constant motion.
2. Gas particles are constantly colliding with each other and the walls of a container. These collisions are elastic; that is, there is no
net loss of energy from the collisions.
3. Gas particles are separated by large distances, with the size of a gas particle tiny compared to the distances that separate them.
4. There are no interactive forces (i.e., attraction or repulsion) between the particles of a gas.
5. The average speed of gas particles is dependent on the temperature of the gas.
Figure 4.1.1 shows a representation of how we mentally picture the gas phase.

Figure 4.1.1 : The Kinetic Theory of Gases. The kinetic theory of gases describes this state of matter as composed of tiny particles
in constant motion with a lot of distance between the particles.
This model of gases explains some of the physical properties of gases. Because most of a gas is empty space, a gas has a low
density and can expand or contract under the appropriate influence. The fact that gas particles are in constant motion means that
two or more gases will always mix, as the particles from the individual gases move and collide with each other.
An ideal gas is a gas that exactly follows the statements of the kinetic theory. Unfortunately, real gases are not ideal. Many gases
deviate slightly from agreeing perfectly with the kinetic theory of gases. However, most gases adhere to the statements so well that
the kinetic theory of gases is well accepted by the scientific community.
The physical behavior of gases is explained by the kinetic theory of gases.
An ideal gas adheres exactly to the kinetic theory of gases.

4.1: Kinetic Molecular Theory- A Model for Gases is shared under a CC BY-NC-SA 3.0 license and was authored, remixed, and/or curated by
Marisa Alviar-Agnew & Henry Agnew.
11.2: Kinetic Molecular Theory- A Model for Gases by Henry Agnew, Marisa Alviar-Agnew is licensed CC BY-NC-SA 3.0. Original
source: https://www.ck12.org/c/chemistry/.

4.1.1 https://chem.libretexts.org/@go/page/423632
4.2: Pressure - The Result of Constant Molecular Collisions
 Learning Objectives
Define pressure.
Learn the units of pressure and how to convert between them.

The kinetic theory of gases indicates that gas particles are always in motion and are colliding with other particles and the walls of
the container holding them. Although collisions with container walls are elastic (i.e., there is no net energy gain or loss because of
the collision), a gas particle does exert a force on the wall during the collision. The accumulation of all these forces distributed over
the area of the walls of the container causes something we call pressure. Pressure (P ) is defined as the force of all the gas
particle/wall collisions divided by the area of the wall:
force
pressure =
area

All gases exert pressure; it is one of the fundamental measurable quantities of this phase of matter. Even our atmosphere exerts
pressure—in this case, the gas is being “held in” by the earth’s gravity, rather than the gas being in a container. The pressure of the
atmosphere is about 14.7 pounds of force for every square inch of surface area: 14.7 lb/in2.
Pressure has a variety of units. The formal, SI-approved unit of pressure is the pascal (Pa), which is defined as 1 N/m2 (one newton
of force over an area of one square meter). However, this is usually too small in magnitude to be useful. A common unit of pressure
is the atmosphere (atm), which was originally defined as the average atmospheric pressure at sea level.
However, “average atmospheric pressure at sea level” is difficult to pinpoint because of atmospheric pressure variations. A more
reliable and common unit is millimeters of mercury (mmHg), which is the amount of pressure exerted by a column of mercury
exactly 1 mm high. An equivalent unit is the torr, which equals 1 mmHg. (The torr is named after Evangelista Torricelli, a
seventeenth-century Italian scientist who invented the mercury barometer.) With these definitions of pressure, the atmosphere unit
is redefined: 1 atm is defined as exactly 760 mmHg, or 760 torr. We thus have the following equivalents:
1 atm=760 mmHg=760 torr
We can use these equivalents as with any equivalence—to perform conversions from one unit to another. Relating these to the
formal SI unit of pressure, 1 atm = 101,325 Pa.

 Example 4.2.1: Pressure Conversion


How many atmospheres are there in 595 torr?

Solution
Solutions to Example 11.3.1
Steps for Problem Solving Unit Conversion

Identify the "given” information and what the problem is asking you Given: 595 torr
to "find." Find: ? atm

List other known quantities. 1 atm = 760 torr

Prepare a concept map.

1 atm
Cancel units and calculate. 595 torr × = 0.783 atm
760 torr

595 torr is less than 760 torr so the final answer should be less than 1
Think about your result.
atm.

4.2.1 https://chem.libretexts.org/@go/page/423633
 Exercise 4.2.1

How many atmospheres are there in 1,022 torr?

Answer
1.345 atm

 Example 4.2.2: Mars

The atmosphere on Mars is largely CO2 at a pressure of 6.01 mmHg. What is this pressure in atmospheres?

Solution
Solutions to Example 11.3.2
Steps for Problem Solving Unit Conversion

Identify the "given” information and what the problem is asking you Given: 6.01mmHg
to "find." Find: ? atm

List other known quantities. 1 atm = 760 mmHg

Prepare a concept map.

1 atm
−3
Cancel units and calculate. 6.01 mmHg × = 0.00791 atm = 7.91 × 10 atm
760 mmHg

6.01 is a very small number relative to 760 mmHg, just like the value
Think about your result.
in atmospheres.

 Exercise 4.2.2
Atmospheric pressure is low in the eye of a hurricane. In a 1979 hurricane in the Pacific Ocean, a pressure of 0.859 atm was
reported inside the eye. What is this pressure in torr?

Answer
652 torr

Summary
Pressure is a force exerted over an area.
Pressure has several common units that can be converted.

4.2: Pressure - The Result of Constant Molecular Collisions is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa
Alviar-Agnew & Henry Agnew.
11.3: Pressure - The Result of Constant Molecular Collisions by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

4.2.2 https://chem.libretexts.org/@go/page/423633
4.3: Boyle’s Law - Pressure and Volume
 Learning Objectives
Learn what is meant by the term gas laws.
Learn and apply Boyle’s Law.

When seventeenth-century scientists began studying the physical properties of gases, they noticed some simple relationships
between some of the measurable properties of the gas. Take pressure (P) and volume (V), for example. Scientists noted that for a
given amount of a gas (usually expressed in units of moles [n]), if the temperature (T) of the gas was kept constant, pressure and
volume were related: as one increases, the other decreases. As one decreases, the other increases. This means that pressure and
volume are inversely related.
There is more to it, however: pressure and volume of a given amount of gas at constant temperature are numerically related. If you
take the pressure value and multiply it by the volume value, the product is a constant for a given amount of gas at a constant
temperature:

P × V = constant at constant n and T

If either volume or pressure changes while amount and temperature stay the same, then the other property must change so that the
product of the two properties still equals that same constant. That is, if the original conditions are labeled P and V and the new
1 1

conditions are labeled P and V , we have


2 2

P1 V1 = constant = P2 V2

where the properties are assumed to be multiplied together. Leaving out the middle part, we have simply

P1 V1 = P2 V2 at constant n and T

This equation is an example of a gas law. A gas law is a simple mathematical formula that allows you to model, or predict, the
behavior of a gas. This particular gas law is called Boyle's Law, after the English scientist Robert Boyle, who first announced it in
1662. Figure 4.3.1 shows two representations of how Boyle’s Law works.

Figure 4.3.1 : Boyle’s Law. A piston having a certain pressure and volume (left piston) will have half the volume when its pressure
is twice as much (right piston). One can also plot P versus V for a given amount of gas at a certain temperature; such a plot will
look like the graph on the right.
Boyle’s Law is an example of a second type of mathematical problem we see in chemistry—one based on a mathematical formula.
Tactics for working with mathematical formulas are different from tactics for working with conversion factors. First, most of the
questions you will have to answer using formulas are word-type questions, so the first step is to identify what quantities are known
and assign them to variables. Second, in most formulas, some mathematical rearrangements (i.e., algebra) must be performed to
solve for an unknown variable. The rule is that to find the value of the unknown variable, you must mathematically isolate the
unknown variable by itself and in the numerator of one side of the equation. Finally, units must be consistent. For example, in
Boyle’s Law there are two pressure variables; they must have the same unit. There are also two volume variables; they also must
have the same unit. In most cases, it won’t matter what the unit is, but the unit must be the same on both sides of the equation.

4.3.1 https://chem.libretexts.org/@go/page/423634
 Example 4.3.1

A sample of gas has an initial pressure of 2.44 atm and an initial volume of 4.01 L. Its pressure changes to 1.93 atm. What is
the new volume if temperature and amount are kept constant?

Solution
Solutions to Example 11.8.1
Steps for Problem Solving

Given: P1 = 2.44 atm and V1 = 4.01 L


Identify the "given" information and what the problem is asking you
P2 = 1.93 atm
to "find."
Find: V2 = ? L

List other known quantities. none

First, rearrange the equation algebraically to solve for V .


2

Plan the problem. P1 × V1


V2 =
P2

Now substitute the known quantities into the equation and solve.

Cancel units and calculate. 2.44 atm × 4.01 L


V2 = = 5.07 L
1.93 atm

We know that pressure and volume are inversely related; as one


decreases, the other increases. Pressure is decreasing (from 2.44 atm
Think about your result. to 1.93 atm), so volume should be increasing to compensate, and it is
(from 4.01 L to 5.07 L). So the answer makes sense based on Boyle’s
Law.

 Exercise 4.3.1

If P1 = 334 torr, V1 = 37.8 mL, and P2 = 102 torr, what is V2?

Answer
124 mL

As mentioned, you can use any units for pressure and volume, but both pressures must be expressed in the same units, and both
volumes must be expressed in the same units.

 Example 4.3.2:
A sample of gas has an initial pressure of 722 torr and an initial volume of 88.8 mL. Its volume changes to 0.663 L. What is the
new pressure?

Solution
Solutions to Example 11.8.2
Steps for Problem Solving

Given: P1 = 722 torr and V1 = 88.8 mL


Identify the "given" information and what the problem is asking you
V2 = 0.633 L
to "find."
Find: P2 = ? torr

List other known quantities. 1 L = 1000 mL to have the same units for volume.

4.3.2 https://chem.libretexts.org/@go/page/423634
Steps for Problem Solving

1. Perform the conversion of the second volume unit from L to mL.


2. Rearrange the equation algebraically to solve for P .
2

Plan the problem.


P1 × V1
P2 =
V2

1.
1000 ml
0.663 L × = 663 ml
1 L

Cancel units and calculate.


2. Substitute the known quantities into the equation and solve.

722 torr × 88.8 mL


P2 = = 96.7 torr
663 mL

When the volume increased, the pressure decreased, which is as


Think about your result.
expected for Boyle’s Law.

 Exercise 4.3.2

If V1 = 456 mL, P1 = 308 torr, and P2 = 1.55 atm, what is V2?

Answer
119 mL

Summary
The behavior of gases can be modeled with gas laws.
Boyle’s Law relates the pressure and volume of a gas at constant temperature and amount.

4.3: Boyle’s Law - Pressure and Volume is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew &
Henry Agnew.
11.4: Boyle’s Law - Pressure and Volume by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

4.3.3 https://chem.libretexts.org/@go/page/423634
4.4: Charles’s Law- Volume and Temperature
 Learning Objectives
Learn and apply Charles's Law.

Everybody enjoys the smell and taste of freshly-baked bread. It is light and fluffy as a result of the action of yeast on sugar. The
yeast converts the sugar to carbon dioxide, which at high temperatures causes the dough to expand. The end result is an enjoyable
treat, especially when covered with melted butter.

Charles's Law
French physicist Jacques Charles (1746-1823) studied the effect of temperature on the volume of a gas at constant pressure.
Charles's Law states that the volume of a given mass of gas varies directly with the absolute temperature of the gas when pressure
is kept constant. The absolute temperature is temperature measured with the Kelvin scale. The Kelvin scale must be used because
zero on the Kelvin scale corresponds to a complete stop of molecular motion.

Figure 4.4.1 : As a container of confined gas is heated, its molecules increase in kinetic energy and push the movable piston
outward, resulting in an increase in volume.
Mathematically, the direct relationship of Charles's Law can be represented by the following equation:
V
=k
T

As with Boyle's Law, k is constant only for a given gas sample. The table below shows temperature and volume data for a set
amount of gas at a constant pressure. The third column is the constant for this particular data set and is always equal to the volume
divided by the Kelvin temperature.
Table 4.4.1 : Temperature-Volume Data
V mL
Temperature (K) Volume (mL) = k ( )
T K

50 20 0.40

100 40 0.40

150 60 0.40

200 80 0.40

300 120 0.40

500 200 0.40

1000 400 0.40

When this data is graphed, the result is a straight line, indicative of a direct relationship, shown in the figure below.

4.4.1 https://chem.libretexts.org/@go/page/423635
Figure 4.4.2 :The volume of a gas increases as the Kelvin temperature increases.
Notice that the line goes exactly toward the origin, meaning that as the absolute temperature of the gas approaches zero, its volume
approaches zero. However, when a gas is brought to extremely cold temperatures, its molecules would eventually condense into the
liquid state before reaching absolute zero. The temperature at which this change into the liquid state occurs varies for different
gases.
Charles's Law can also be used to compare changing conditions for a gas. Now we use V and T to stand for the initial volume
1 1

and temperature of a gas, while V and T stand for the final volume and temperature. The mathematical relationship of Charles's
2 2

Law becomes:
V1 V2
=
T1 T2

This equation can be used to calculate any one of the four quantities if the other three are known. The direct relationship will only
hold if the temperatures are expressed in Kelvin. Temperatures in Celsius will not work. Recall the relationship that
C + 273 .
o
K=

 Example 4.4.1:

A balloon is filled to a volume of 2.20 L at a temperature of 22 o


C . The balloon is then heated to a temperature of 71 o
C . Find
the new volume of the balloon.

Solution
Solutions to Example 11.5.1
Steps for Problem Solving

Given:
V1 = 2.20 L and
Identify the "given" information and what the problem is asking you o
T1 = 22 C = 295 K
to "find." o
T2 = 71 C = 344 K

Find: V2 = ? L

List other known quantities. The temperatures have first been converted to Kelvin.

First, rearrange the equation algebraically to solve for V .2

Plan the problem. V1 × T2


V2 =
T1

Now substitute the known quantities into the equation and solve.

Cancel units and calculate. 2.20 L × 344 K


V2 = = 2.57 L
295 K

4.4.2 https://chem.libretexts.org/@go/page/423635
Steps for Problem Solving

The volume increases as the temperature increases. The result has


Think about your result.
three significant figures.

 Exercise 4.4.1

If V1 = 3.77 L and T1 = 255 K, what is V2 if T2 = 123 K?

Answer
1.82 L

 Example 4.4.2:

A sample of a gas has an initial volume of 34.8 L and an initial temperature of −67°C. What must be the temperature of the gas
for its volume to be 25.0 L?

Solution
Solutions to Example 11.5.2
Steps for Problem Solving

Given:
Identify the "given" information and what the problem is asking you Given:T1 = -27oC and V1 = 34.8 L
to "find." V2 = 25.0 L
Find: T2 = ? K

List other known quantities. K = -27oC + 273

1. Convert the initial temperature to Kelvin


2. Rearrange the equation algebraically to solve for T .2

Plan the problem.


V2 × T1
T2 =
V1

1. −67°C + 273 = 206 K


2. Substitute the known quantities into the equation and solve.
Cancel units and calculate.
25.0 L × 206 K
T2 = = 148 K
34.8 L

This is also equal to −125°C. As temperature decreases, volume


Think about your result.
decreases—which it does in this example.

 Exercise 4.4.2
If V1 = 623 mL, T1 = 255°C, and V2 = 277 mL, what is T2?

Answer
235 K, or −38°C

Summary
Charles’s Law relates the volume and temperature of a gas at constant pressure and amount.

4.4: Charles’s Law- Volume and Temperature is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-
Agnew & Henry Agnew.

4.4.3 https://chem.libretexts.org/@go/page/423635
11.5: Charles’s Law- Volume and Temperature by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

4.4.4 https://chem.libretexts.org/@go/page/423635
4.5: Gay-Lussac's Law- Temperature and Pressure
 Learning Objectives
Explain Gay-Lussac's Law.

Propane tanks are widely used with barbeque grills. But it's not fun to find out halfway through grilling that you have run out of
gas. You can buy gauges that measure the pressure inside the tank to see how much is left. The gauge measures pressure and will
register a higher pressure on a hot day than it will on a cold day. So you need to take the air temperature into account when you
decide whether or not to refill the tank before your next cook-out.

Gay-Lussac's Law
When the temperature of a sample of gas in a rigid container is increased, the pressure of the gas increases as well. The increase in
kinetic energy results in the molecules of gas striking the walls of the container with more force, resulting in a greater pressure. The
French chemist Joseph Gay-Lussac (1778-1850) discovered the relationship between the pressure of a gas and its absolute
temperature. Gay-Lussac's Law states that the pressure of a given mass of gas varies directly with the absolute temperature of the
gas, when the volume is kept constant. Gay-Lussac's Law is very similar to Charles's Law, with the only difference being the type
of container. Whereas the container in a Charles's Law experiment is flexible, it is rigid in a Gay-Lussac's Law experiment.

Figure 4.5.1 : Joseph Gay-Lussac.


The mathematical expressions for Gay-Lussac's Law are likewise similar to those of Charles's Law:
P P1 P2
and =
T T1 T2

A graph of pressure vs. temperature also illustrates a direct relationship. As a gas is cooled at constant volume, its pressure
continually decreases until the gas condenses to a liquid.

 Example 4.5.1
The gas in an aerosol can is under a pressure of 3.00 atm at a temperature of 25 C. It is dangerous to dispose of an aerosol can
o

by incineration. What would the pressure in the aerosol can be at a temperature of 845 C? o

Solution
Solutions to Example 11.10.1
Steps for Problem Solving

Given:
P1 = 3.00 atm
Identify the "given" information and what the problem is asking you o
T1 = 25 C = 298 K
to "find." o
T2 = 845 C = 1118 K

Find: P 2 =? atm

List other known quantities. The temperatures have first been converted to Kelvin.

First, rearrange the equation algebraically to solve for P .


2

Plan the problem. P1 × T2


P2 =
T1

4.5.1 https://chem.libretexts.org/@go/page/423636
Steps for Problem Solving

Now substitute the known quantities into the equation and solve.

Calculate. 3.00 atm × 1118 K


P2 = = 11.3 atm
298 K

The pressure increases dramatically due to a large increase in


Think about your result.
temperature.

Summary
Pressure and temperature at constant volume are directly proportional.

4.5: Gay-Lussac's Law- Temperature and Pressure is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-
Agnew & Henry Agnew.
11.11: Gay-Lussac's Law- Temperature and Pressure by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

4.5.2 https://chem.libretexts.org/@go/page/423636
4.6: Avogadro’s Law- Volume and Moles
A flat tire is not very useful. It does not cushion the rim of the wheel and creates a very uncomfortable ride. When air is added to
the tire, the pressure increases as more molecules of gas are forced into the rigid tire. How much air should be put into a tire
depends on the pressure rating for that tire. Too little pressure and the tire will not hold its shape. Too much pressure and the tire
could burst.

Avogadro's Law
You have learned about Avogadro's hypothesis: equal volumes of any gas at the same temperature and pressure contain the same
number of molecules. It follows that the volume of a gas is directly proportional to the number of moles of gas present in the
sample. Avogadro's Law states that the volume of a gas is directly proportional to the number of moles (or number of particles) of
gas when the temperature and pressure are held constant. The mathematical expression of Avogadro's Law is:

V = k×n

or
V1 V2
=
n1 n2

where n is the number of moles of gas and k is a constant. Avogadro's Law is in evidence whenever you blow up a balloon. The
volume of the balloon increases as you add moles of gas to the balloon by blowing it up.
If the container holding the gas is rigid rather than flexible, pressure can be substituted for volume in Avogadro's Law. Adding gas
to a rigid container makes the pressure increase.

 Example 4.6.1
A balloon has been filled to a volume of 1.90 L with 0.0920 mol of helium gas. If 0.0210 mol of additional helium is added to
the balloon while the temperature and pressure are held constant, what is the new volume of the balloon?

Solution
Solutions to Example 11.11.1
Steps for Problem Solving

Given:
Identify the "given" information and what the problem is asking you V1 = 1.90 L

to "find." n1 = 0.0920 mol

Find: V 2 =? L

Note that the final number of moles has to be calculated by adding the
List other known quantities. original number of moles to the moles of added helium.
n2 = 0.0920 + 0.0210 = 0.1130 mol

First, rearrange the equation algebraically to solve for V . 2

Plan the problem. V1 × n2


V2 =
n1

Now substitute the known quantities into the equation and solve.

Calculate. 1.90 L × 0.1130 mol


V2 = = 2.33 L
0.0920 mol

Since a relatively small amount of additional helium was added to the


Think about your result.
balloon, its volume increases slightly.

4.6.1 https://chem.libretexts.org/@go/page/423637
 Exercise 4.6.1

A 12.8 L volume of gas contains .000498 moles of oxygen gas. At constant temperature and pressure, what volume does
.0000136 moles of the gas fill?

Answer
0.350 L

Summary
Calculations for relationships between volume and number of moles of a gas can be performed using Avogadro's Law.

4.6: Avogadro’s Law- Volume and Moles is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew &
Henry Agnew.
11.7: Avogadro’s Law- Volume and Moles by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

4.6.2 https://chem.libretexts.org/@go/page/423637
4.7: The Ideal Gas Law- Pressure, Volume, Temperature, and Moles
 Learning Objectives
Explain the Ideal Gas Law.

There are a number of chemical reactions that require ammonia. In order to carry out the reaction efficiently, we need to know how
much ammonia we have for stoichiometric purposes. Using gas laws, we can determine the number of moles present in the tank if
we know the volume, temperature, and pressure of the system.

Ideal Gas Law


The Combined Gas Law shows that the pressure of a gas is inversely proportional to volume and directly proportional to
temperature. Avogadro's Law shows that volume or pressure is directly proportional to the number of moles of gas. Putting these
together leaves us with the following equation:
P1 × V1 P2 × V2
=
T1 × n1 T2 × n2

(P × V )
As with the other gas laws, we can also say that is equal to a constant. The constant can be evaluated provided that the
(T × n)

gas being described is considered to be ideal.


The Ideal Gas Law is a single equation which relates the pressure, volume, temperature, and number of moles of an ideal gas. If
we substitute in the variable R for the constant, the equation becomes:
P ×V
=R
T ×n

The Ideal Gas Law is conveniently rearranged to look this way, with the multiplication signs omitted:

P V = nRT

The variable R in the equation is called the ideal gas constant.

Evaluating the Ideal Gas Constant


The value of R , the ideal gas constant, depends on the units chosen for pressure, temperature, and volume in the ideal gas equation.
It is necessary to use Kelvin for the temperature and it is conventional to use the SI unit of liters for the volume. However, pressure
is commonly measured in one of three units: kPa , atm, or mm Hg. Therefore, R can have three different values.
We will demonstrate how R is calculated when the pressure is measured in kPa . The volume of 1.00 mol of any gas at STP
(Standard temperature, 273.15 K and pressure, 1 atm) is measured to be 22.414 L. We can substitute 101.325 kPa for pressure,
22.414 L for volume, and 273.15 K for temperature into the ideal gas equation and solve for R .

PV
R =
nT

101.325 kPa × 22.414 L


=
1.000 mol × 273.15 K

= 8.314 kPa ⋅ L/K ⋅ mol

This is the value of R that is to be used in the ideal gas equation when the pressure is given in kPa . The table below shows a
summary of this and the other possible values of R . It is important to choose the correct value of R to use for a given problem.
Table 4.7.1 : Values of the Ideal Gas Constant
Unit of P Unit of V Unit of n Unit of T Value and Unit of R

kPa L mol K 8.314 J/K ⋅ mol

atm L mol K 0.08206 L ⋅ atm/K ⋅ mol

mm Hg L mol K 62.36 L ⋅ mm Hg/K ⋅ mol

4.7.1 https://chem.libretexts.org/@go/page/423638
Notice that the unit for R when the pressure is in kPa has been changed to J/K ⋅ mol. A kilopascal multiplied by a liter is equal to
the SI unit for energy, a joule (J).

 Example 4.7.1 Oxygen Gas


What volume is occupied by 3.76 g of oxygen gas at a pressure of 88.4 kPa and a temperature of 19 o
C ? Assume the oxygen is
ideal.

Solution
Solutions to Example 11.5.1
Steps for Problem Solving

Given:
P = 88.4 kPa
Identify the "given" information and what the problem is asking you o
T = 19 C = 292 K
to "find."
Mass O = 3.76 g
2

Find: V = ? L
O = 32.00 g/mol
List other known quantities. 2

R = 8.314 J/K ⋅ mol

1. First, determine the number of moles of O2 from the given mass


and the molar mass.
Plan the problem. 2. Then, rearrange the equation algebraically to solve for V
nRT
V =
P

1.
1 mol O
2
3.76 g × = 0.1175 mol O
2
32.00 g O
2

Calculate. 2. Now substitute the known quantities into the equation and solve.

0.1175 mol × 8.314 J/K ⋅ mol × 292 K


nRT
V = = = 3.23 L O
2
P
88.4 kPa

The number of moles of oxygen is far less than one mole, so the
volume should be fairly small compared to molar volume
(22.4 L/mol) since the pressure and temperature are reasonably close
Think about your result.
to standard. The result has three significant figures because of the
values for T and P . Since a joule (J) = kPa ⋅ L , the units cancel out
correctly, leaving a volume in liters.

 Example 4.7.2: Argon Gas


A 4.22 mol sample of Ar has a pressure of 1.21 atm and a temperature of 34°C. What is its volume?

Solution
Solutions to Example 11.5.2
Steps for Problem Solving

4.7.2 https://chem.libretexts.org/@go/page/423638
Steps for Problem Solving

Given:
n = 4.22 mol
Identify the "given" information and what the problem is asking you
P = 1.21 atm
to "find."
T = 34°C
Find: V = ? L

List other known quantities. none

1. The first step is to convert temperature to Kelvin.


2. Then, rearrange the equation algebraically to solve for V
Plan the problem.
nRT
V =
P

1. 34 + 273 = 307 K
2. Now substitute the known quantities into the equation and solve.

L. atm

Calculate. (4.22 mol )(0.08205 )(307 K )


mol. K
V =
1.21 atm

= 87.9 L

The number of moles of Ar is large so the expected volume should


Think about your result.
also be large.

 Exercise 4.7.1

A 0.0997 mol sample of O2 has a pressure of 0.692 atm and a temperature of 333 K. What is its volume?

Answer
3.94 L

 Exercise 4.7.2

For a 0.00554 mol sample of H2, P = 23.44 torr and T = 557 K. What is its volume?

Answer
8.21 L

Summary
The Ideal Gas Law is a single equation which relates the pressure, volume, temperature, and number of moles of an ideal gas.

4.7: The Ideal Gas Law- Pressure, Volume, Temperature, and Moles is shared under a CK-12 license and was authored, remixed, and/or curated
by Marisa Alviar-Agnew & Henry Agnew.
11.8: The Ideal Gas Law- Pressure, Volume, Temperature, and Moles by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original
source: https://www.ck12.org/c/chemistry/.

4.7.3 https://chem.libretexts.org/@go/page/423638
4.8: Mixtures of Gases - Why Deep-Sea Divers Breathe a Mixture of Helium and
Oxygen

 Learning Objectives
Explain Dalton's Law of Partial Pressures.

The atmosphere of Venus is markedly different from that of Earth. The gases in the Venusian atmosphere are 96.5% carbon dioxide
and 3% nitrogen. The atmospheric pressure on Venus is roughly 92 times that of Earth, so the amount of nitrogen on Venus would
contribute a pressure well over 2700 mm Hg. And there is no oxygen present, so we couldn't breathe there. Not that we would
want to go to Venus, as the surface temperature is usually over 460 C. o

Dalton's Law of Partial Pressures


Gas pressure results from collisions between gas particles and the inside walls of their container. If more gas is added to a rigid
container, the gas pressure increases. The identities of the two gases do not matter. John Dalton, the English chemist who proposed
the atomic theory, also studied mixtures of gases. He found that each gas in a mixture exerts a pressure independently of every
other gas in the mixture. For example, our atmosphere is composed of about 78% nitrogen and 21% oxygen, with smaller amounts
of several other gases making up the rest. Since nitrogen makes up 78% of the gas particles in a given sample of air, it exerts 78%
of the pressure. If the overall atmospheric pressure is 1.00 atm, then the pressure of just the nitrogen in the air is 0.78 atm. The
pressure of the oxygen in the air is 0.21 atm.
The partial pressure of a gas is the contribution that gas makes to the total pressure when the gas is part of a mixture. The partial
pressure of nitrogen is represented by P . Dalton's Law of Partial Pressures states that the total pressure of a mixture of gases is
N2

equal to the sum of all of the partial pressures of the component gases. Dalton's Law can be expressed with the following equation:

Ptotal = P1 + P2 + P3 + ⋯

The figure below shows two gases that are in separate, equal-sized containers at the same temperature and pressure. Each exerts a
different pressure, P and P , reflective of the number of particles in the container. On the right, the two gases are combined into
1 2

the same container, with no volume change. The total pressure of the gas mixture is equal to the sum of the individual pressures. If
P = 300 mm Hg and P = 500 mm Hg , then P
1 2 = 800 mm Hg.
total

Figure 4.8.1 : Dalton's Law states that the pressure of a gas mixture is equal to the partial pressures of the combining gases.

Collecting Gases Over Water


You need to do a lab experiment where hydrogen gas is generated. In order to calculate the yield of gas, you have to know the
pressure inside the tube where the gas is collected. But how can you get a barometer in there? Very simple: you don't. All you need

4.8.1 https://chem.libretexts.org/@go/page/423639
is the atmospheric pressure in the room. As the gas pushes out the water, it is pushing against the atmosphere, so the pressure inside
is equal to the pressure outside.

Gas Collection by Water Displacement


Gases that are produced in laboratory experiments are often collected by a technique called water displacement (Figure 4.8.2). A
bottle is filled with water and placed upside-down in a pan of water. The reaction flask is fitted with rubber tubing, which is then
fed under the bottle of water. As the gas is produced in the reaction flask, it exits through the rubber tubing and displaces the water
in the bottle. When the bottle is full of the gas, it can be sealed with a lid.

Figure 4.8.2 : A gas produced in a chemical reaction can be collected by water displacement.
Because the gas is collected over water, it is not pure, but is mixed with vapor from the evaporation of the water. Dalton's Law can
be used to calculate the amount of the desired gas by subtracting the contribution of the water vapor.

PTotal = Pg + PH O
2

where P is the pressure of the desired gas, which can be solved for:
g

Pg = PT otal − PH O
2

In order to solve a problem, it is necessary to know the vapor pressure of water at the temperature of the reaction (see table below).
The sample problem illustrates the use of Dalton's Law when a gas is collected over water.
Table 4.8.1 : Vapor Pressure of Water (mm Hg) at Selected Temperatures ( o
C)

0 5 10 15 20 25 30 35 40 45 50 55 60

4.58 6.54 9.21 12.79 17.54 23.76 31.82 42.18 55.32 71.88 92.51 118.04 149.38

 Example 14.14.1

A certain experiment generates 2.58 L of hydrogen gas, which is collected over water. The temperature is 20 C
o
and the
atmospheric pressure is 98.60 kPa. Find the volume that the dry hydrogen would occupy at STP.

Solution

Step 1: List the known quantities and plan the problem.

Known
VTotal = 2.58 L
o
T = 20 C = 293 K

PTotal = 98.60 kPa = 739.7 mm Hg

Unknown
VH2 at STP =? L
The atmospheric pressure is converted from kPa to mm Hg in order to match units with the table. The sum of the pressures of
the hydrogen and the water vapor is equal to the atmospheric pressure. The pressure of the hydrogen is found by subtraction.

4.8.2 https://chem.libretexts.org/@go/page/423639
Then, the volume of the gas at STP can be calculated by using the combined gas law.

Step 2: Solve.
PH = PTotal − PH O
2 2

= 739, 7 mm Hg − 17.54 mm Hg

= 722.2 mm Hg

Now the combined gas law is used, solving for V , the volume of hydrogen at STP.
2

P1 × V1 × T2
V2 =
P2 × T1

722.2 mm Hg × 2.58 L × 273 K


=
760 mm Hg × 293 K

= 2.28 L H
2

Step 3: Think about your result.


If the hydrogen gas were to be collected at STP and without the presence of the water vapor, its volume would be 2.28 L. This
is less than the actual collected volume because some of that is water vapor. The conversion using STP is useful for
stoichiometry purposes.

Summary
Dalton's Law of Partial Pressures states that the total pressure in a system is equal to the sum of the partial pressures of the
gases present.
The vapor pressure due to water in a sample can be corrected for, in order to get the true value for the pressure of the gas.

4.8: Mixtures of Gases - Why Deep-Sea Divers Breathe a Mixture of Helium and Oxygen is shared under a CK-12 license and was authored,
remixed, and/or curated by Marisa Alviar-Agnew & Henry Agnew.
11.9: Mixtures of Gases - Why Deep-Sea Divers Breathe a Mixture of Helium and Oxygen by Henry Agnew, Marisa Alviar-Agnew is
licensed CK-12. Original source: https://www.ck12.org/c/chemistry/.

4.8.3 https://chem.libretexts.org/@go/page/423639
4.9: Interactions between Molecules
In the winter, many people find the snow and ice beautiful; they enjoy getting out to ski or ice-skate. When the snow melts,
however, the roads get very sloppy and messy. Some people look forward to spring, when the ice and snow are gone and the
weather is warmer. All of these events and factors are dependent on the melting point of a solid and the freezing point of a liquid.

Melting Point
Solids are similar to liquids in that both are condensed states, with particles that are far closer together than those of a gas.
However, while liquids are fluid, solids are not. The particles of most solids are packed tightly together in an orderly arrangement.
The motion of individual atoms, ions, or molecules in a solid is restricted to vibrational motion about a fixed point. Solids are
almost completely incompressible and are the densest of the three states of matter.
As a solid is heated, its particles vibrate more rapidly as it absorbs kinetic energy. Eventually, the organization of the particles
within the solid structure begins to break down and the solid starts to melt. The melting point is the temperature at which a solid
changes into a liquid. At its melting point, the disruptive vibrations of the particles of the solid overcome the attractive forces
operating within the solid. As with boiling points, the melting point of a solid is dependent on the strength of those attractive
forces. Sodium chloride (NaCl) is an ionic compound that consists of a multitude of strong ionic bonds. Sodium chloride melts at
801 C. Ice (solid H O ) is a molecular compound of molecules that are held together by hydrogen bonds. Though hydrogen bonds
o
2

are the strongest of the intermolecular forces, the strength of hydrogen bonds is much less than that of ionic bonds. The melting
point of ice is 0 °C.

Figure 4.9.1 : Melting ice cubes illustrate the process of this phase transition. (Public Domain; Moussa).
The melting point of a solid is the same as the freezing point of the liquid. At that temperature, the solid and liquid states of the
substance are in equilibrium. For water, this equilibrium occurs at 0 C. o

H O (s) ⇌ H O (l)
2 2

We tend to think of solids as those materials that are solid at room temperature. However, all materials have melting points of some
sort. Gases become solids at extremely low temperatures, and liquids will also become solid if the temperature is low enough. The
table below gives the melting points of some common materials.
Table 4.9.1 : Melting Points of Common Materials
Materials Melting Point (°C)

Hydrogen -259

Oxygen -219

Diethyl ether -116

Ethanol -114

Water 0

Pure silver 961

4.9.1 https://chem.libretexts.org/@go/page/423640
Materials Melting Point (°C)

Pure gold 1063

Iron 1538

Summary
The melting point is the temperature at which a solid changes into a liquid. Intermolecular forces have a strong influence on
melting point.

4.9: Interactions between Molecules is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew &
Henry Agnew.
12.1: Interactions between Molecules by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

4.9.2 https://chem.libretexts.org/@go/page/423640
4.10: Properties of Liquids and Solids
 Learning Objectives
Describe the solid and liquid phases.

Solids and liquids are collectively called condensed phases because their particles are in virtual contact. The two states share little
else, however.

Solids
In the solid state, the individual particles of a substance are in fixed positions with respect to each other because there is not enough
thermal energy to overcome the intermolecular interactions between the particles. As a result, solids have a definite shape and
volume. Most solids are hard, but some (like waxes) are relatively soft. Many solids composed of ions can also be quite brittle.

Figure 4.10.1 : A crystalline arrangement of quartz crystal cluster. Some large crystals look the way they do because of the regular
arrangement of atoms (ions) in their crystal structure. (Source: Wikipedia.)
Solids usually have their constituent particles arranged in a regular, three-dimensional array of alternating positive and negative
ions called a crystal. The effect of this regular arrangement of particles is sometimes visible macroscopically, as shown in Figure
4.10.1. Some solids, especially those composed of large molecules, cannot easily organize their particles in such regular crystals

and exist as amorphous (literally, “without form”) solids. Glass is one example of an amorphous solid.

Liquids
If the particles of a substance have enough energy to partially overcome intermolecular interactions, then the particles can move
about each other while remaining in contact. This describes the liquid state. In a liquid, the particles are still in close contact, so
liquids have a definite volume. However, because the particles can move about each other rather freely, a liquid has no definite
shape and takes a shape dictated by its container.

Figure 4.10.2 : The formation of a spherical droplet of liquid water minimizes the surface area, which is the natural result of surface
tension in liquids. (Source: Wikipedia.)

Gases
If the particles of a substance have enough energy to completely overcome intermolecular interactions, then the particles can
separate from each other and move about randomly in space. Like liquids, gases have no definite shape, but unlike solids and
liquids, gases have no definite volume either.

4.10.1 https://chem.libretexts.org/@go/page/423641
Figure 4.10.3 : A Representation of the Solid, Liquid, and Gas States. A solid has definite volume and shape, a liquid has a definite
volume but no definite shape, and a gas has neither a definite volume nor shape.
The change from solid to liquid usually does not significantly change the volume of a substance. However, the change from a liquid
to a gas significantly increases the volume of a substance, by a factor of 1,000 or more. Figure 4.10.3 shows the differences among
solids, liquids, and gases at the molecular level, while Table 4.10.1 lists the different characteristics of these states.
Table 4.10.1 : Characteristics of the Three States of Matter
Characteristic Solid Liquid Gas

shape definite indefinite indefinite

volume definite definite indefinite

relative intermolecular interaction


strong moderate weak
strength

relative particle positions in contact and fixed in place in contact but not fixed not in contact, random positions

 Example 4.10.1

What state or states of matter does each statement describe?


a. This state has a definite volume.
b. This state has no definite shape.
c. This state allows the individual particles to move about while remaining in contact.

Solution
a. This statement describes either the liquid state or the solid state.
b. This statement describes either the liquid state or the gas state.
c. This statement describes the liquid state.

 Exercise 4.10.1

What state or states of matter does each statement describe?


a. This state has individual particles in a fixed position with regard to each other.
b. This state has individual particles far apart from each other in space.
c. This state has indefinite shape.

Answer a
solid
Answer b
gas
Answer c
liquid or gas

4.10.2 https://chem.libretexts.org/@go/page/423641
 Looking Closer: Water, the Most Important Liquid

Earth is the only known body in our solar system that has liquid water existing freely on its surface; life on Earth would not be
possible without the presence of liquid water.
Water has several properties that make it a unique substance among substances. It is an excellent solvent; it dissolves many
other substances and allows those substances to react when in solution. In fact, water is sometimes called the universal solvent
because of this ability. Water has unusually high melting and boiling points (0°C and 100°C, respectively) for such a small
molecule. The boiling points for similar-sized molecules, such as methane (BP = −162°C) and ammonia (BP = −33°C), are
more than 100° lower. Though a liquid at normal temperatures, water molecules experience a relatively strong intermolecular
interaction that allows them to maintain the liquid phase at higher temperatures than expected.
Unlike most substances, the solid form of water is less dense than its liquid form, which allows ice to float on water. In colder
weather, lakes and rivers freeze from the top, allowing animals and plants to continue to live underneath. Water also requires
an unusually large amount of energy to change temperature. While 100 J of energy will change the temperature of 1 g of Fe by
230°C, this same amount of energy will change the temperature of 1 g of H2O by only 100°C. Thus, water changes its
temperature slowly as heat is added or removed. This has a major impact on weather, as storm systems like hurricanes can be
impacted by the amount of heat that ocean water can store.
Water’s influence on the world around us is affected by these properties. Isn’t it fascinating that such a small molecule can
have such a big impact?

Key Takeaway
Solids and liquids are phases that have their own unique properties.

4.10: Properties of Liquids and Solids is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew &
Henry Agnew.
12.2: Properties of Liquids and Solids by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

4.10.3 https://chem.libretexts.org/@go/page/423641
4.11: Intermolecular Forces in Action- Surface Tension and Viscosity
 Learning Objectives
Explain the how the surface tension of a liquid relates to intermolecular forces.

The next time you are by a still body of water, take a close look at what is scooting along on the surface. You may see insects
seemingly floating on top of the water. These creatures are known by a variety of names including water skaters, water striders,
pond skaters, and other equally descriptive names. They take advantage of a property called surface tension to stay above the water
and not sink. The force they exert downward is less than the forces exerted among the water molecules on the surface of the pond,
so the insect does not penetrate beneath the surface of the water.

Surface Tension
Molecules within a liquid are pulled equally in all directions by intermolecular forces. However, molecules at the surface are pulled
downwards and sideways by other liquid molecules, but not upwards away from the surface. The overall effect is that the surface
molecules are pulled into the liquid, creating a surface that is tightened like a film (Figure 4.11.1A). The surface tension of a
liquid is a measure of the elastic force in the liquid's surface. Liquids that have strong intermolecular forces, like the hydrogen
bonding in water, exhibit the greatest surface tension. Surface tension allows objects that are denser than water, such as the paper
clip shown in B in the figure below, to nonetheless float on its surface. It is also responsible for the beading up of water droplets on
a freshly waxed car, because there are no attractions between the polar water molecules and the nonpolar wax.

Figure 4.11.1 : (A) Molecules at the surface of a liquid are pulled downwards into the liquid, creating a tightened surface. (B)
Surface tension allows a paper clip to float on water's surface.
Other liquids, such as diethyl ether, do not demonstrate strong surface tension interactions. The intermolecular forces for the ether
are the relatively weak dipole-dipole interactions that do not draw the molecules together as tightly as hydrogen bonds would.

Summary
The surface tension of a liquid is a measure of the elastic force in the liquid's surface.
Liquids with strong intermolecular forces have higher surface tensions than liquids with weaker forces.

4.11: Intermolecular Forces in Action- Surface Tension and Viscosity is shared under a CK-12 license and was authored, remixed, and/or curated
by Marisa Alviar-Agnew & Henry Agnew.
12.3: Intermolecular Forces in Action- Surface Tension and Viscosity by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original
source: https://www.ck12.org/c/chemistry/.

4.11.1 https://chem.libretexts.org/@go/page/423642
4.12: Evaporation and Condensation
 Learning Objectives
Explain how intermolecular forces affect rates of vaporization, evaporation, and condensation.

On the roof of the house in the picture below is a device known as a "swamp cooler". This piece of equipment traces its origin back
to the ancient Egyptians who hung wet blankets across the doors of their homes. As the warm air passed through the blankets,
water would evaporate and cool the air. The royalty went one step further and had servants fan wet cloths over jugs of water to get
more evaporation and cooling.

Figure 4.12.1 : A home with a swamp cooler attached to the roof.


The origin of the term "swamp cooler" is not known - they certainly don't work in a swamp. Best conditions for cooling include a
high temperature (over 80 F) and a low humidity (preferably less than 30%). These coolers work well in desert areas, but don't
o

provide any cooling in the humid areas of the country.

Evaporation
A puddle of water left undisturbed eventually disappears. The liquid molecules escape into the gas phase, becoming water vapor.
Vaporization is the process in which a liquid is converted to a gas. Evaporation is the conversion of a liquid to its vapor below the
boiling temperature of the liquid. If the water is instead kept in a closed container, the water vapor molecules do not have a chance
to escape into the surroundings and so the water level does not change. As some water molecules become vapor, an equal number
of water vapor molecules condense back into the liquid state. Condensation is the change of state from a gas to a liquid.

Figure 4.12.2 : Evaporation (A) and condensation (B).


In order for a liquid molecule to escape into the gas state, the molecule must have enough kinetic energy to overcome the
intermolecular attractive forces in the liquid. Recall that a given liquid sample will have molecules with a wide range of kinetic
energies. Liquid molecules that have this certain threshold kinetic energy escape the surface and become vapor. As a result, the
liquid molecules that remain now have lower kinetic energy. As evaporation occurs, the temperature of the remaining liquid
decreases. You have observed the effects of evaporative cooling. On a hot day, the water molecules in your perspiration absorb
body heat and evaporate from the surface of your skin. The evaporating process leaves the remaining perspiration cooler, which in
turn absorbs more heat from your body.
A given liquid will evaporate more quickly when it is heated. This is because the heating process results in a greater fraction of the
liquid's molecules having the necessary kinetic energy to escape the surface of the liquid. The figure below shows the kinetic
energy distribution of liquid molecules at two temperatures. The numbers of molecules that have the required kinetic energy to

4.12.1 https://chem.libretexts.org/@go/page/423643
evaporate are shown in the shaded area under the curve at the right. The higher temperature liquid (T 2) has more molecules that are
capable of escaping into the vapor phase than the lower temperature liquid (T ) . 1

Figure 4.12.3 : Kinetic energy distribution curves for a liquid at two temperatures T1 and T2 . The shaded area represents the
molecules with enough kinetic energy to escape the liquid and become vapor.
At 29,029 feet (8848 m), Mount Everest in the Himalayan range on the border between China and Nepal is the highest point on the
earth. Its altitude presents many practical problems to climbers. The oxygen content of the air is much lower than at sea level,
making it necessary to bring oxygen tanks along (although a few climbers have reached the peak without oxygen). One other
problem is that of boiling water for cooking food. Although water boils at 100 C at sea level, the boiling point on top of Mount
o

Everest is only about 70 C. This difference makes it very difficult to get a decent cup of tea (which definitely frustrated some of
o

the British climbers).

Boiling
As a liquid is heated, the average kinetic energy of its particles increases. The rate of evaporation increases as more and more
molecules are able to escape the liquid's surface into the vapor phase. Eventually a point is reached when the molecules all
throughout the liquid have enough kinetic energy to vaporize. At this point the liquid begins to boil. The boiling point is the
temperature at which the vapor pressure of a liquid is equal to the external pressure. The figure below illustrates the boiling of
liquid.

Figure 4.12.4 : Comparison between evaporation and boiling.


In the picture on the left, the liquid is below its boiling point, yet some of the liquid evaporates. On the right, the temperature has
been increased until bubbles begin to form in the body of the liquid. When the vapor pressure inside the bubble is equal to the
external atmospheric pressure, the bubbles rise to the surface of the liquid and burst. The temperature at which this process occurs
is the boiling point of the liquid.
The normal boiling point is the temperature at which the vapor pressure of the liquid is equal to standard pressure. Because
atmospheric pressure can change based on location, the boiling point of a liquid changes with the external pressure. The normal
boiling point is a constant because it is defined relative to the standard atmospheric pressure of 760 mm Hg (or 1 atm or
101.3 kPa).

4.12.2 https://chem.libretexts.org/@go/page/423643
Figure 4.12.5 : Influence of altitude on the boiling point of water.

Summary
The boiling point is the temperature at which the vapor pressure of a liquid is equal to the external pressure.
As the altitude increases, the boiling point decreases.
Evaporation is the conversion of a liquid to its vapor below the boiling temperature of the liquid.
Condensation is the change of state from a gas to a liquid.
As the temperature increases, the rate of evaporation increases.

4.12: Evaporation and Condensation is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew &
Henry Agnew.
12.4: Evaporation and Condensation by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

4.12.3 https://chem.libretexts.org/@go/page/423643
4.13: Melting, Freezing, and Sublimation
 Learning Objectives
Define melting, freezing, and sublimation.

Depending on the surrounding conditions, normal matter usually exists as one of three phases: solid, liquid, or gas.
A phase change is a physical process in which a substance goes from one phase to another. Usually the change occurs when adding
or removing heat at a particular temperature, known as the melting point or the boiling point of the substance. The melting point is
the temperature at which the substance goes from a solid to a liquid (or from a liquid to a solid). The boiling point is the
temperature at which a substance goes from a liquid to a gas (or from a gas to a liquid). The nature of the phase change depends on
the direction of the heat transfer. Heat going into a substance changes it from a solid to a liquid, or a liquid to a gas. Removing heat
from a substance changes a gas to a liquid, or a liquid to a solid.
Two key points are worth emphasizing. First, at a substance’s melting point or boiling point, two phases can exist simultaneously.
Take water (H2O) as an example. On the Celsius scale, H2O has a melting point of 0°C and a boiling point of 100°C. At 0°C, both
the solid and liquid phases of H2O can coexist. However, if heat is added, some of the solid H2O will melt and turn into liquid H2O.
If heat is removed, the opposite happens: some of the liquid H2O turns into solid H2O. A similar process can occur at 100°C:
adding heat increases the amount of gaseous H2O, while removing heat increases the amount of liquid H2O (Figure 4.13.1).

Figure 4.13.1 : The Boiling Point of Water. Nucleate boiling of water over a kitchen stove burner. (Source: Wikipedia). Water is a
good substance to use as an example because many people are already familiar with it. Other substances have melting points and
boiling points as well.
Second, the temperature of a substance does not change as the substance goes from one phase to another. In other words, phase
changes are isothermal (isothermal means “constant temperature”). Again, consider H2O as an example. Solid water (ice) can exist
at 0°C. If heat is added to ice at 0°C, some of the solid changes phase to make liquid, which is also at 0°C. Remember, the solid and
liquid phases of H2O can coexist at 0°C. Only after all of the solid has melted into liquid does the addition of heat change the
temperature of the substance.
For each phase change of a substance, there is a characteristic quantity of heat needed to perform the phase change per gram (or per
mole) of material. The heat of fusion (ΔHfus) is the amount of heat per gram (or per mole) required for a phase change that occurs
at the melting point. The heat of vaporization (ΔHvap) is the amount of heat per gram (or per mole) required for a phase change that
occurs at the boiling point. If you know the total number of grams or moles of material, you can use the ΔHfus or the ΔHvap to
determine the total heat being transferred for melting or solidification using these expressions:
heat = n × ΔHf us (4.13.1)

where n is the number of moles and ΔH f us is expressed in energy/mole or


heat = m × ΔHf us (4.13.2)

where m is the mass in grams and ΔH f us is expressed in energy/gram.


For the boiling or condensation, use these expressions:
heat = n × ΔHvap (4.13.3)

where n is the number of moles) and ΔH vap is expressed in energy/mole or

4.13.1 https://chem.libretexts.org/@go/page/423644
heat = m × ΔHvap (4.13.4)

where m is the mass in grams and ΔH vap is expressed in energy/gram.


Remember that a phase change depends on the direction of the heat transfer. If heat transfers in, solids become liquids, and liquids
become solids at the melting and boiling points, respectively. If heat transfers out, liquids solidify, and gases condense into liquids.

 Example 4.13.1

How much heat is necessary to melt 55.8 g of ice (solid H2O) at 0°C? The heat of fusion of H2O is 79.9 cal/g.

Solution
We can use the relationship between heat and the heat of fusion (Eq. 4.13.1b) to determine how many joules of heat are needed
to melt this ice:
heat = m × ΔHf us

79.9 cal
= (55.8 g ) ( )
g

= 4, 460 cal

 Exercise 4.13.1

How much heat is necessary to vaporize 685 g of H2O at 100°C? The heat of vaporization of H2O is 540 cal/g.

Table 4.13.1 lists the heats of fusion and vaporization for some common substances. Note the units on these quantities; when you
use these values in problem solving, make sure that the other variables in your calculation are expressed in units consistent with the
units in the specific heats, or the heats of fusion and vaporization.
Table 4.13.1 : Heats of Fusion and Vaporization for Selected Substances
Substance ΔHfus (cal/g) ΔHvap (cal/g)

aluminum (Al) 94.0 2,602

gold (Au) 15.3 409

iron (Fe) 63.2 1,504

water (H2O) 79.9 540

sodium chloride (NaCl) 123.5 691

ethanol (C2H5OH) 45.2 200.3

benzene (C6H6) 30.4 94.1

Looking Closer: Sublimation


There is also a phase change where a solid goes directly to a gas:
solid→ gas (4.13.5)

This phase change is called sublimation. Each substance has a characteristic heat of sublimation associated with this process. For
example, the heat of sublimation (ΔHsub) of H2O is 620 cal/g.
We encounter sublimation in several ways. You may already be familiar with dry ice, which is simply solid carbon dioxide (CO2).
At −78.5°C (−109°F), solid carbon dioxide sublimes, changing directly from the solid phase to the gas phase:

−78.5 C

C O2 (s) −−−−−→ C O2 (g) (4.13.6)

4.13.2 https://chem.libretexts.org/@go/page/423644
Solid carbon dioxide is called dry ice because it does not pass through the liquid phase. Instead, it goes directly to the gas phase.
(Carbon dioxide can exist as liquid but only under high pressure.) Dry ice has many practical uses, including the long-term
preservation of medical samples.
Even at temperatures below 0°C, solid H2O will slowly sublime. For example, a thin layer of snow or frost on the ground may
slowly disappear as the solid H2O sublimes, even though the outside temperature may be below the freezing point of water.
Similarly, ice cubes in a freezer may get smaller over time. Although frozen, the solid water slowly sublimes, redepositing on the
colder cooling elements of the freezer, which necessitates periodic defrosting (frost-free freezers minimize this redeposition).
Lowering the temperature in a freezer will reduce the need to defrost as often.
Under similar circumstances, water will also sublime from frozen foods (e.g., meats or vegetables), giving them an unattractive,
mottled appearance called freezer burn. It is not really a “burn,” and the food has not necessarily gone bad, although it looks
unappetizing. Freezer burn can be minimized by lowering a freezer’s temperature and by wrapping foods tightly so water does not
have any space to sublime into.

Melting Point
Solids are similar to liquids in that both are condensed states, with particles that are far closer together than those of a gas.
However, while liquids are fluid, solids are not. The particles of most solids are packed tightly together in an orderly arrangement.
The motion of individual atoms, ions, or molecules in a solid is restricted to vibrational motion about a fixed point. Solids are
almost completely incompressible and are the most dense of the three states of matter.
As a solid is heated, its particles vibrate more rapidly as the solid absorbs kinetic energy. Eventually, the organization of the
particles within the solid structure begins to break down and the solid starts to melt. The melting point is the temperature at which
a solid changes into a liquid. At its melting point, the disruptive vibrations of the particles of the solid overcome the attractive
forces operating within the solid. As with boiling points, the melting point of a solid is dependent on the strength of those attractive
forces. Sodium chloride (NaCl) is an ionic compound that consists of a multitude of strong ionic bonds. Sodium chloride melts at
801 C. Ice (solid H O ) is a molecular compound composed of molecules that are held together by hydrogen bonds. Though
o

hydrogen bonds are the strongest of the intermolecular forces, the strength of hydrogen bonds is much less than that of ionic bonds.
The melting point of ice is 0 C.
o

The melting point of a solid is the same as the freezing point of the liquid. At that temperature, the solid and liquid states of the
substance are in equilibrium. For water, this equilibrium occurs at 0 C. o

H O (s) ⇌ H O (l)
2 2

We tend to think of solids as those materials that are solid at room temperature. However, all materials have melting points of some
sort. Gases become solids at extremely low temperatures, and liquids will also become solid if the temperature is low enough. The
table below gives the melting points of some common materials.
Table 4.13.2 : Melting Points of Common Materials
Materials Melting Point (ºC)

Hydrogen -259

Oxygen -219

Diethyl ether -116

Ethanol -114

Water 0

Pure silver 961

Pure gold 1063

Iron 1538

4.13.3 https://chem.libretexts.org/@go/page/423644
 Exercise 4.13.2
a. Explain what happens when heat flows into or out of a substance at its melting point or boiling point.
b. How does the amount of heat required for a phase change relate to the mass of the substance?

Answer a
The energy goes into changing the phase, not the temperature.

Answer b
The amount of heat is a constant per gram of substance.

Summary
There is an energy change associated with any phase change.
Sublimation is the change of state from a solid to a gas, without passing through the liquid state.
Deposition is the change of state from a gas to a solid.
Carbon dioxide is an example of a material that easily undergoes sublimation.
The melting point is the temperature at which a solid changes into a liquid.
Intermolecular forces have a strong influence on melting point.

4.13: Melting, Freezing, and Sublimation is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew &
Henry Agnew.
12.5: Melting, Freezing, and Sublimation by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

4.13.4 https://chem.libretexts.org/@go/page/423644
4.14: INTERMOLECULAR FORCES- DISPERSION, DIPOLE–DIPOLE,
HYDROGEN BONDING, AND ION-DIPOLE
 LEARNING OBJECTIVES
To describe the intermolecular forces in liquids.

The properties of liquids are intermediate between those of gases and solids, but are more similar to solids. In contrast to intramolecular
forces, such as the covalent bonds that hold atoms together in molecules and polyatomic ions, intermolecular forces hold molecules together
in a liquid or solid. Intermolecular forces are generally much weaker than covalent bonds. For example, it requires 927 kJ to overcome the
intramolecular forces and break both O–H bonds in 1 mol of water, but it takes only about 41 kJ to overcome the intermolecular attractions
and convert 1 mol of liquid water to water vapor at 100°C. (Despite this seemingly low value, the intermolecular forces in liquid water are
among the strongest such forces known!) Given the large difference in the strengths of intra- and intermolecular forces, changes between the
solid, liquid, and gaseous states almost invariably occur for molecular substances without breaking covalent bonds.

The properties of liquids are intermediate between those of gases and solids, but are more similar to
solids.

Intermolecular forces determine bulk properties, such as the melting points of solids and the boiling points of liquids. Liquids boil when the
molecules have enough thermal energy to overcome the intermolecular attractive forces that hold them together, thereby forming bubbles of
vapor within the liquid. Similarly, solids melt when the molecules acquire enough thermal energy to overcome the intermolecular forces that
lock them into place in the solid.
Intermolecular forces are electrostatic in nature; that is, they arise from the interaction between positively and negatively charged species.
Like covalent and ionic bonds, intermolecular interactions are the sum of both attractive and repulsive components. Because electrostatic
interactions fall off rapidly with increasing distance between molecules, intermolecular interactions are most important for solids and
liquids, where the molecules are close together. These interactions become important for gases only at very high pressures, where they are
responsible for the observed deviations from the ideal gas law at high pressures.
In this section, we explicitly consider three kinds of intermolecular interactions. There are two additional types of electrostatic interaction
that you are already familiar with: the ion–ion interactions that are responsible for ionic bonding, and the ion–dipole interactions that occur
when ionic substances dissolve in a polar substance such as water. The first two are often described collectively as van der Waals forces.

DIPOLE–DIPOLE INTERACTIONS
Polar covalent bonds behave as if the bonded atoms have localized fractional charges that are equal but opposite (i.e., the two bonded atoms
generate a dipole). If the structure of a molecule is such that the individual bond dipoles do not cancel one another, then the molecule has a
net dipole moment. Molecules with net dipole moments tend to align themselves so that the positive end of one dipole is near the negative
end of another and vice versa, as shown in Figure 4.14.1a .

Figure 4.14.1: Attractive and Repulsive Dipole–Dipole Interactions. (a and b) Molecular orientations in which the positive end of one
dipole (δ+) is near the negative end of another (δ−) (and vice versa) produce attractive interactions. (c and d) Molecular orientations that
juxtapose the positive or negative ends of the dipoles on adjacent molecules produce repulsive interactions. (CC BY-SA-NC; anonymous)
These arrangements are more stable than arrangements in which two positive or two negative ends are adjacent (Figure 4.14.1c ). Hence
dipole–dipole interactions, such as those in Figure 4.14.1b , are attractive intermolecular interactions, whereas those in Figure 4.14.1d are
repulsive intermolecular interactions. Because molecules in a liquid move freely and continuously, molecules always experience both
attractive and repulsive dipole–dipole interactions simultaneously, as shown in Figure 4.14.2 . On average, however, the attractive
interactions dominate.

4.14.1 https://chem.libretexts.org/@go/page/423645
Figure 4.14.2: Both attractive and repulsive dipole–dipole interactions occur in a liquid sample with many molecules. (CC BY-SA-NC;
anonymous)
The green arrows pointing towards each other represent attraction. The gray arrows pointing away from each other represent repulsion
Because each end of a dipole possesses only a fraction of the charge of an electron, dipole–dipole interactions are substantially weaker than
the interactions between two ions, each of which has a charge of at least ±1, or between a dipole and an ion, in which one of the species has
at least a full positive or negative charge. In addition, the attractive interaction between dipoles falls off much more rapidly with increasing
distance than do the ion–ion interactions. Recall that the attractive energy between two ions is proportional to 1/r, where r is the distance
between the ions. Doubling the distance (r → 2r) decreases the attractive energy by one-half. In contrast, the energy of the interaction of
two dipoles is proportional to 1/r3, so doubling the distance between the dipoles decreases the strength of the interaction by 23, or 8-fold.
Thus a substance such as HCl , which is partially held together by dipole–dipole interactions, is a gas at room temperature and 1 atm
pressure. Conversely, NaCl , which is held together by interionic interactions, is a high-melting-point solid. Within a series of compounds of
similar molar mass, the strength of the intermolecular interactions increases as the dipole moment of the molecules increases, as shown in
Table 4.14.1 .
Table 4.14.1: Relationships Between the Dipole Moment and the Boiling Point for Organic Compounds of Similar Molar Mass
Compound Molar Mass (g/mol) Dipole Moment (D) Boiling Point (K)
C3H6 (cyclopropane) 42 0 240
CH3OCH3 (dimethyl ether) 46 1.30 248
CH3CN (acetonitrile) 41 3.9 355

The attractive energy between two ions is proportional to 1/r, whereas the attractive energy between
two dipoles is proportional to 1/r6.

Dipole Intermolecular Force

Video Discussing Dipole Intermolecular Forces. Source: Dipole Intermolecular Force, YouTube(opens in new window) [youtu.be]

4.14.2 https://chem.libretexts.org/@go/page/423645
 EXAMPLE 4.14.1

Arrange ethyl methyl ether (CH3OCH2CH3), 2-methylpropane [isobutane, (CH3)2CHCH3], and acetone (CH3COCH3) in order of
increasing boiling points. Their structures are as follows:

Given: compounds.
Asked for: order of increasing boiling points.

Strategy:
Compare the molar masses and the polarities of the compounds. Compounds with higher molar masses and that are polar will have the
highest boiling points.

Solution:
The three compounds have essentially the same molar mass (58–60 g/mol), so we must look at differences in polarity to predict the
strength of the intermolecular dipole–dipole interactions and thus the boiling points of the compounds.
The first compound, 2-methylpropane, contains only C–H bonds, which are not very polar because C and H have similar
electronegativities. It should therefore have a very small (but nonzero) dipole moment and a very low boiling point.
Ethyl methyl ether has a structure similar to H2O; it contains two polar C–O single bonds oriented at about a 109° angle to each other,
in addition to relatively nonpolar C–H bonds. As a result, the C–O bond dipoles partially reinforce one another and generate a
significant dipole moment that should give a moderately high boiling point.
Acetone contains a polar C=O double bond oriented at about 120° to two methyl groups with nonpolar C–H bonds. The C–O bond
dipole therefore corresponds to the molecular dipole, which should result in both a rather large dipole moment and a high boiling point.
Thus we predict the following order of boiling points:

2-methylpropane < ethyl methyl ether < acetone


This result is in good agreement with the actual data: 2-methylpropane, boiling point = −11.7°C, and the dipole moment (μ) = 0.13 D;
methyl ethyl ether, boiling point = 7.4°C and μ = 1.17 D; acetone, boiling point = 56.1°C and μ = 2.88 D.

 EXERCISE 4.14.1

Arrange carbon tetrafluoride (CF4), ethyl methyl sulfide (CH3SC2H5), dimethyl sulfoxide [(CH3)2S=O], and 2-methylbutane
[isopentane, (CH3)2CHCH2CH3] in order of decreasing boiling points.

Answer
dimethyl sulfoxide (boiling point = 189.9°C) > ethyl methyl sulfide (boiling point = 67°C) > 2-methylbutane (boiling point =
27.8°C) > carbon tetrafluoride (boiling point = −128°C)

4.14.3 https://chem.libretexts.org/@go/page/423645
LONDON DISPERSION FORCES
Thus far, we have considered only interactions between polar molecules. Other factors must be considered to explain why many nonpolar
molecules, such as bromine, benzene, and hexane, are liquids at room temperature; why others, such as iodine and naphthalene, are solids.
Even the noble gases can be liquefied or solidified at low temperatures, high pressures, or both (Table 4.14.2 ).
What kind of attractive forces can exist between nonpolar molecules or atoms? This question was answered by Fritz London (1900–1954), a
German physicist who later worked in the United States. In 1930, London proposed that temporary fluctuations in the electron distributions
within atoms and nonpolar molecules could result in the formation of short-lived instantaneous dipole moments, which produce attractive
forces called London dispersion forces between otherwise nonpolar substances.
Table 4.14.2: Normal Melting and Boiling Points of Some Elements and Nonpolar Compounds
Substance Molar Mass (g/mol) Melting Point (°C) Boiling Point (°C)
Ar 40 −189.4 −185.9
Xe 131 −111.8 −108.1
N2 28 −210 −195.8
O2 32 −218.8 −183.0
F2 38 −219.7 −188.1
I2 254 113.7 184.4
CH4 16 −182.5 −161.5

Consider a pair of adjacent He atoms, for example. On average, the two electrons in each He atom are uniformly distributed around the
nucleus. Because the electrons are in constant motion, however, their distribution in one atom is likely to be asymmetrical at any given
instant, resulting in an instantaneous dipole moment. As shown in part (a) in Figure 4.14.3 , the instantaneous dipole moment on one atom
can interact with the electrons in an adjacent atom, pulling them toward the positive end of the instantaneous dipole or repelling them from
the negative end. The net effect is that the first atom causes the temporary formation of a dipole, called an induced dipole, in the second.
Interactions between these temporary dipoles cause atoms to be attracted to one another. These attractive interactions are weak and fall off
rapidly with increasing distance. London was able to show with quantum mechanics that the attractive energy between molecules due to
temporary dipole–induced dipole interactions falls off as 1/r6. Doubling the distance therefore decreases the attractive energy by 26, or 64-
fold.

Figure 4.14.3: Instantaneous Dipole Moments. The formation of an instantaneous dipole moment on one He atom (a) or an H2 molecule (b)
results in the formation of an induced dipole on an adjacent atom or molecule.
Instantaneous dipole–induced dipole interactions between nonpolar molecules can produce intermolecular attractions just as they produce
interatomic attractions in monatomic substances like Xe. This effect, illustrated for two H2 molecules in part (b) in Figure 4.14.3 , tends to

4.14.4 https://chem.libretexts.org/@go/page/423645
become more pronounced as atomic and molecular masses increase (Table 4.14.2 ). For example, Xe boils at −108.1°C, whereas He boils at
−269°C. The reason for this trend is that the strength of London dispersion forces is related to the ease with which the electron distribution
in a given atom can be perturbed. In small atoms such as He, the two 1s electrons are held close to the nucleus in a very small volume, and
electron–electron repulsions are strong enough to prevent significant asymmetry in their distribution. In larger atoms such as Xe, however,
the outer electrons are much less strongly attracted to the nucleus because of filled intervening shells. As a result, it is relatively easy to
temporarily deform the electron distribution to generate an instantaneous or induced dipole. The ease of deformation of the electron
distribution in an atom or molecule is called its polarizability. Because the electron distribution is more easily perturbed in large, heavy
species than in small, light species, we say that heavier substances tend to be much more polarizable than lighter ones.

For similar substances, London dispersion forces get stronger with increasing molecular size.

The polarizability of a substance also determines how it interacts with ions and species that possess permanent dipoles. Thus, London
dispersion forces are responsible for the general trend toward higher boiling points with increased molecular mass and greater surface area
in a homologous series of compounds, such as the alkanes (part (a) in Figure 4.14.4 ). The strengths of London dispersion forces also
depend significantly on molecular shape because shape determines how much of one molecule can interact with its neighboring molecules
at any given time. For example, part (b) in Figure 4.14.4 shows 2,2-dimethylpropane (neopentane) and n-pentane, both of which have the
empirical formula C5H12. Neopentane is almost spherical, with a small surface area for intermolecular interactions, whereas n-pentane has
an extended conformation that enables it to come into close contact with other n-pentane molecules. As a result, the boiling point of
neopentane (9.5°C) is more than 25°C lower than the boiling point of n-pentane (36.1°C).

Figure 4.14.4: Mass and Surface Area Affect the Strength of London Dispersion Forces. (a) In this series of four simple alkanes, larger
molecules have stronger London forces between them than smaller molecules do, and consequently have higher boiling points. (b) Linear n-
pentane molecules have a larger surface area and stronger intermolecular forces than spherical neopentane molecules. As a result,
neopentane is a gas at room temperature, whereas n-pentane is a volatile liquid.
All molecules, whether polar or nonpolar, are attracted to one another by London dispersion forces in addition to any other attractive forces
that may be present. In general, however, dipole–dipole interactions in small polar molecules are significantly stronger than London
dispersion forces, so the former predominate.

Dispersion Intermolecular Force

Video Discussing London/Dispersion Intermolecular Forces. Source: Dispersion Intermolecular Force, YouTube(opens in new window)
[youtu.be]

4.14.5 https://chem.libretexts.org/@go/page/423645
 EXAMPLE 4.14.2

Arrange n-butane, propane, 2-methylpropane [isobutene, (CH3)2CHCH3], and n-pentane in order of increasing boiling points.
Given: compounds
Asked for: order of increasing boiling points

Strategy:
Determine the intermolecular forces in the compounds, and then arrange the compounds according to the strength of those forces. The
substance with the weakest forces will have the lowest boiling point.

Solution:
The four compounds are alkanes and nonpolar, so London dispersion forces are the only important intermolecular forces. These forces
are generally stronger with increasing molecular mass, so propane should have the lowest boiling point and n-pentane should have the
highest, with the two butane isomers falling in between. Of the two butane isomers, 2-methylpropane is more compact, and n-butane
has the more extended shape. Consequently, we expect intermolecular interactions for n-butane to be stronger due to its larger surface
area, resulting in a higher boiling point. The overall order is thus as follows, with actual boiling points in parentheses: propane
(−42.1°C) < 2-methylpropane (−11.7°C) < n-butane (−0.5°C) < n-pentane (36.1°C).

 EXERCISE 4.14.2

Arrange GeH4, SiCl4, SiH4, CH4, and GeCl4 in order of decreasing boiling points.

Answer
GeCl4 (87°C) > SiCl4 (57.6°C) > GeH4 (−88.5°C) > SiH4 (−111.8°C) > CH4 (−161°C)

HYDROGEN BONDS
Molecules with hydrogen atoms bonded to electronegative atoms such as O, N, and F (and to a much lesser extent, Cl and S) tend to exhibit
unusually strong intermolecular interactions. These result in much higher boiling points than are observed for substances in which London
dispersion forces dominate, as illustrated for the covalent hydrides of elements of groups 14–17 in Figure 4.14.5 . Methane and its heavier
congeners in group 14 form a series whose boiling points increase smoothly with increasing molar mass. This is the expected trend in
nonpolar molecules, for which London dispersion forces are the exclusive intermolecular forces. In contrast, the hydrides of the lightest
members of groups 15–17 have boiling points that are more than 100°C greater than predicted on the basis of their molar masses. The effect
is most dramatic for water: if we extend the straight line connecting the points for H2Te and H2Se to the line for period 2, we obtain an
estimated boiling point of −130°C for water! Imagine the implications for life on Earth if water boiled at −130°C rather than 100°C.

4.14.6 https://chem.libretexts.org/@go/page/423645
Figure 4.14.5: The Effects of Hydrogen Bonding on Boiling Points. These plots of the boiling points of the covalent hydrides of the
elements of groups 14–17 show that the boiling points of the lightest members of each series for which hydrogen bonding is possible (HF,
NH3, and H2O) are anomalously high for compounds with such low molecular masses.
Group 14 is in purple, group 15 is in green, group 16 is red, and group 17 is blue. Graph of boiling point against period.
Why do strong intermolecular forces produce such anomalously high boiling points and other unusual properties, such as high enthalpies of
vaporization and high melting points? The answer lies in the highly polar nature of the bonds between hydrogen and very electronegative
elements such as O, N, and F. The large difference in electronegativity results in a large partial positive charge on hydrogen and a
correspondingly large partial negative charge on the O, N, or F atom. Consequently, H–O, H–N, and H–F bonds have very large bond
dipoles that can interact strongly with one another. Because a hydrogen atom is so small, these dipoles can also approach one another more
closely than most other dipoles. The combination of large bond dipoles and short dipole–dipole distances results in very strong dipole–
dipole interactions called hydrogen bonds, as shown for ice in Figure 4.14.6 . A hydrogen bond is usually indicated by a dotted line between
the hydrogen atom attached to O, N, or F (the hydrogen bond donor) and the atom that has the lone pair of electrons (the hydrogen bond
acceptor). Because each water molecule contains two hydrogen atoms and two lone pairs, a tetrahedral arrangement maximizes the number
of hydrogen bonds that can be formed. In the structure of ice, each oxygen atom is surrounded by a distorted tetrahedron of hydrogen atoms
that form bridges to the oxygen atoms of adjacent water molecules. The bridging hydrogen atoms are not equidistant from the two oxygen
atoms they connect, however. Instead, each hydrogen atom is 101 pm from one oxygen and 174 pm from the other. In contrast, each oxygen
atom is bonded to two H atoms at the shorter distance and two at the longer distance, corresponding to two O–H covalent bonds and two
O⋅⋅⋅H hydrogen bonds from adjacent water molecules, respectively. The resulting open, cagelike structure of ice means that the solid is
actually slightly less dense than the liquid, which explains why ice floats on water, rather than sinks.

Figure 4.14.6: The Hydrogen-Bonded Structure of Ice

4.14.7 https://chem.libretexts.org/@go/page/423645
Each water molecule accepts two hydrogen bonds from two other water molecules and donates two hydrogen atoms to form hydrogen
bonds with two more water molecules, producing an open, cage like structure. The structure of liquid water is very similar, but in the liquid,
the hydrogen bonds are continually broken and formed because of rapid molecular motion.

Hydrogen bond formation requires both a hydrogen bond donor and a hydrogen bond acceptor.

Because ice is less dense than liquid water, rivers, lakes, and oceans freeze from the top down. In fact, the ice forms a protective surface
layer that insulates the rest of the water, allowing fish and other organisms to survive in the lower levels of a frozen lake or sea. If ice were
denser than the liquid, the ice formed at the surface in cold weather would sink as fast as it formed. Bodies of water would freeze from the
bottom up, which would be lethal for most aquatic creatures. The expansion of water when freezing also explains why automobile or boat
engines must be protected by “antifreeze” and why unprotected pipes in houses break if they are allowed to freeze.

Hydrogen Bonding Intermolecular Force

Video Discussing Hydrogen Bonding Intermolecular Forces. Source: Hydrogen Bonding Intermolecular Force, YouTube(opens in new
window) [youtu.be]

 EXAMPLE 4.14.3

Considering CH3OH, C2H6, Xe, and (CH3)3N, which can form hydrogen bonds with themselves? Draw the hydrogen-bonded
structures.
Given: compounds
Asked for: formation of hydrogen bonds and structure

Strategy:
A. Identify the compounds with a hydrogen atom attached to O, N, or F. These are likely to be able to act as hydrogen bond donors.
B. Of the compounds that can act as hydrogen bond donors, identify those that also contain lone pairs of electrons, which allow them
to be hydrogen bond acceptors. If a substance is both a hydrogen donor and a hydrogen bond acceptor, draw a structure showing the
hydrogen bonding.

Solution:
A. Of the species listed, xenon (Xe), ethane (C2H6), and trimethylamine [(CH3)3N] do not contain a hydrogen atom attached to O, N, or
F; hence they cannot act as hydrogen bond donors.
B. The one compound that can act as a hydrogen bond donor, methanol (CH3OH), contains both a hydrogen atom attached to O
(making it a hydrogen bond donor) and two lone pairs of electrons on O (making it a hydrogen bond acceptor); methanol can thus form
hydrogen bonds by acting as either a hydrogen bond donor or a hydrogen bond acceptor. The hydrogen-bonded structure of methanol is
as follows:

4.14.8 https://chem.libretexts.org/@go/page/423645
 EXERCISE 4.14.3

Considering CH3CO2H, (CH3)3N, NH3, and CH3F, which can form hydrogen bonds with themselves? Draw the hydrogen-bonded
structures.

Answer
CH3CO2H and NH3;

Hydrogen bonding in ammonia between nitrogen and hydrogen. hydrogen bonding in acetic acid is between oxygen and hydrogen.

Although hydrogen bonds are significantly weaker than covalent bonds, with typical dissociation energies of only 15–25 kJ/mol, they have
a significant influence on the physical properties of a compound. Compounds such as HF can form only two hydrogen bonds at a time as
can, on average, pure liquid NH3. Consequently, even though their molecular masses are similar to that of water, their boiling points are
significantly lower than the boiling point of water, which forms four hydrogen bonds at a time.

 EXAMPLE 4.14.4: BUCKYBALLS

Arrange C60 (buckminsterfullerene, which has a cage structure), NaCl, He, Ar, and N2O in order of increasing boiling points.
Given: compounds.
Asked for: order of increasing boiling points.

Strategy:
Identify the intermolecular forces in each compound and then arrange the compounds according to the strength of those forces. The
substance with the weakest forces will have the lowest boiling point.

Solution
Electrostatic interactions are strongest for an ionic compound, so we expect NaCl to have the highest boiling point. To predict the
relative boiling points of the other compounds, we must consider their polarity (for dipole–dipole interactions), their ability to form
hydrogen bonds, and their molar mass (for London dispersion forces). Helium is nonpolar and by far the lightest, so it should have the
lowest boiling point. Argon and N2O have very similar molar masses (40 and 44 g/mol, respectively), but N2O is polar while Ar is not.
Consequently, N2O should have a higher boiling point. A C60 molecule is nonpolar, but its molar mass is 720 g/mol, much greater than
that of Ar or N2O. Because the boiling points of nonpolar substances increase rapidly with molecular mass, C60 should boil at a higher
temperature than the other nonionic substances. The predicted order is thus as follows, with actual boiling points in parentheses:
He (−269°C) < Ar (−185.7°C) < N2O (−88.5°C) < C60 (>280°C) < NaCl (1465°C).

4.14.9 https://chem.libretexts.org/@go/page/423645
 EXERCISE 4.14.4

Arrange 2,4-dimethylheptane, Ne, CS2, Cl2, and KBr in order of decreasing boiling points.

Answer
KBr (1435°C) > 2,4-dimethylheptane (132.9°C) > CS2 (46.6°C) > Cl2 (−34.6°C) > Ne (−246°C)

 EXAMPLE 4.14.5

Identify the most significant intermolecular force in each substance.


a. C3H8
b. CH3OH
c. H2S

Solution
a. Although C–H bonds are polar, they are only minimally polar. The most significant intermolecular force for this substance would be
dispersion forces.
b. This molecule has an H atom bonded to an O atom, so it will experience hydrogen bonding.
c. Although this molecule does not experience hydrogen bonding, the Lewis electron dot diagram and VSEPR indicate that it is bent,
so it has a permanent dipole. The most significant force in this substance is dipole-dipole interaction.

 EXERCISE 4.14.6

Identify the most significant intermolecular force in each substance.


a. HF
b. HCl

Answer a
hydrogen bonding
Answer b
dipole-dipole interactions

SUMMARY
Intermolecular forces are electrostatic in nature and include van der Waals forces and hydrogen bonds. Molecules in liquids are held to other
molecules by intermolecular interactions, which are weaker than the intramolecular interactions that hold the atoms together within
molecules and polyatomic ions. Transitions between the solid and liquid, or the liquid and gas phases, are due to changes in intermolecular
interactions, but do not affect intramolecular interactions. The three major types of intermolecular interactions are dipole–dipole
interactions, London dispersion forces (these two are often referred to collectively as van der Waals forces), and hydrogen bonds. Dipole–
dipole interactions arise from the electrostatic interactions of the positive and negative ends of molecules with permanent dipole moments;
their strength is proportional to the magnitude of the dipole moment and to 1/r3, where r is the distance between dipoles. London
dispersion forces are due to the formation of instantaneous dipole moments in polar or nonpolar molecules as a result of short-lived
fluctuations of electron charge distribution, which in turn cause the temporary formation of an induced dipole in adjacent molecules; their
energy falls off as 1/r6. Larger atoms tend to be more polarizable than smaller ones, because their outer electrons are less tightly bound and
are therefore more easily perturbed. Hydrogen bonds are especially strong dipole–dipole interactions between molecules that have
hydrogen bonded to a highly electronegative atom, such as O, N, or F. The resulting partially positively charged H atom on one molecule
(the hydrogen bond donor) can interact strongly with a lone pair of electrons of a partially negatively charged O, N, or F atom on adjacent
molecules (the hydrogen bond acceptor). Because of strong O⋅⋅⋅H hydrogen bonding between water molecules, water has an unusually high
boiling point, and ice has an open, cage like structure that is less dense than liquid water.

4.14: Intermolecular Forces- Dispersion, Dipole–Dipole, Hydrogen Bonding, and Ion-Dipole is shared under a CC BY-NC-SA 3.0 license and was
authored, remixed, and/or curated by LibreTexts.
11.2: Intermolecular Forces is licensed CC BY-NC-SA 3.0.

4.14.10 https://chem.libretexts.org/@go/page/423645
4.15: Water - A Remarkable Molecule
 Learning Objectives
Interpret the unique properties of water in terms of a phase diagram.

Earth is the only known body in our solar system that has liquid water existing freely on its surface; life on Earth would not be
possible without the presence of liquid water. Water has several properties that make it a unique substance among substances. It is
an excellent solvent; it dissolves many other substances and allows those substances to react when in solution. In fact, water is
sometimes called the universal solvent because of this ability. Water has unusually high melting and boiling points (0°C and 100°C,
respectively) for such a small molecule. The boiling points for similar-sized molecules, such as methane (BP = −162°C) and
ammonia (BP = −33°C), are more than 100° lower. Though a liquid at normal temperatures, water molecules experience a
relatively strong intermolecular interaction that allows them to maintain the liquid phase at higher temperatures than expected.
Unlike most substances, the solid form of water is less dense than its liquid form, which allows ice to float on water. In colder
weather, lakes and rivers freeze from the top, allowing animals and plants to continue to live underneath. Water also requires an
unusually large amount of energy to change temperature. While 100 J of energy will change the temperature of 1 g of Fe by 230°C,
this same amount of energy will change the temperature of 1 g of H2O by only 100°C. Thus, water changes its temperature slowly
as heat is added or removed. This has a major impact on weather, as storm systems like hurricanes can be impacted by the amount
of heat that ocean water can store. Water’s influence on the world around us is affected by these properties. Isn’t it fascinating that
such a small molecule can have such a big impact?

Phase Diagram for Water


Water is a unique substance in many ways. One of these special properties is the fact that solid water (ice) is less dense than liquid
water just above the freezing point. The phase diagram for water is shown in the figure below.

Figure 4.15.1 : Phase diagram for water.


Notice one key difference between the general phase diagram and the phase diagram for water. In water's diagram, the slope of the
line between the solid and liquid states is negative rather than positive. The reason for this is that water is an unusual substance, as
its solid state is less dense than the liquid state. Ice floats in liquid water. Therefore, a pressure change has the opposite effect on
those two phases. If ice is relatively near its melting point, it can be changed into liquid water by the application of pressure. The
water molecules are actually closer together in the liquid phase than they are in the solid phase.
Refer again to water's phase diagram (figure above). Notice point E , labeled the critical point. What does that mean? At
373.99 C, particles of water in the gas phase are moving very, very rapidly. At any temperature higher than that, the gas phase
o

cannot be made to liquefy, no matter how much pressure is applied to the gas. The critical pressure (P ) is the pressure that must
C

be applied to the gas at the critical temperature in order to turn it into a liquid. For water, the critical pressure is very high,
217.75 atm. The critical point is the intersection point of the critical temperature and the critical pressure.

4.15.1 https://chem.libretexts.org/@go/page/423646
Summary
Solid water is less dense than liquid water just above the freezing point.
The critical temperature (T ) of a substance is the highest temperature at which the substance can possibly exist as a liquid.
C

The critical pressure (P ) is the pressure that must be applied to the gas at the critical temperature in order to turn it into a
C

liquid.
The critical point is the intersection point of the critical temperature and the critical pressure.

4.15: Water - A Remarkable Molecule is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew &
Henry Agnew.
12.8: Water - A Remarkable Molecule by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

4.15.2 https://chem.libretexts.org/@go/page/423646
CHAPTER OVERVIEW

5: Module 5 - Solution Chemistry


5.1: Solutions - Homogeneous Mixtures
5.2: Solutions of Solids Dissolved in Water- How to Make Rock Candy
5.3: Specifying Solution Concentration- Mass Percent
5.4: Specifying Solution Concentration- Molarity
5.5: Solution Dilution
5.6: Solution Stoichiometry
5.7: Electrolytes
5.8: Acids- Properties and Examples
5.9: Bases- Properties and Examples
5.10: Molecular Definitions of Acids and Bases
5.11: Reactions of Acids and Bases
5.12: Strong and Weak Acids and Bases
5.13: Acid–Base Titration
5.14: Water - Acid and Base in One
5.15: The pH and pOH Scales - Ways to Express Acidity and Basicity
5.16: Buffers- Solutions that Resist pH Change

5: Module 5 - Solution Chemistry is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.

1
5.1: SOLUTIONS - HOMOGENEOUS MIXTURES
 LEARNING OBJECTIVES
Learn terminology involving solutions.
Explain the significance of the statement "like dissolves like."
Explain why certain substances dissolve in other substances.

The major component of a solution is called the solvent. The minor component of a solution is called the solute. By major and minor we
mean whichever component has the greater or lesser presence by mass or by moles. Sometimes this becomes confusing, especially with
substances with very different molar masses. However, here we will confine the discussion to solutions for which the major component and
the minor component are obvious.

Figure 5.1.1: Making a saline water solution by dissolving table salt (NaCl) in water. The salt is the solute and the water the solvent. (CC-
BY-SA 3.0; Chris 73).
Solutions exist for every possible phase of the solute and the solvent. Salt water, for example, is a solution of solid NaCl in liquid water,
while air is a solution of a gaseous solute (O ) in a gaseous solvent (N ). In all cases, however, the overall phase of the solution is the same
2 2

phase as the solvent. Table 5.1.1 lists some common types of solutions, with examples of each.
Table 5.1.1: Types of Solutions
Solvent Phase Solute Phase Example
gas gas air
liquid gas carbonated beverages
liquid liquid ethanol (C2H5OH) in H2O (alcoholic beverages)
liquid solid salt water
solid gas H2 gas absorbed by Pd metal
solid liquid Hg(ℓ) in dental fillings
solid solid steel alloys

 EXAMPLE 5.1.1: SUGAR AND WATER


A solution is made by dissolving 1.00 g of sucrose (C 12
H
22
O
11
) in 100.0 g of liquid water. Identify the solvent and solute in the
resulting solution.

Solution
Either by mass or by moles, the obvious minor component is sucrose, so it is the solute. Water—the majority component—is the
solvent. The fact that the resulting solution is the same phase as water also suggests that water is the solvent.

 EXERCISE 5.1.1

A solution is made by dissolving 3.33 g of HCl(g) in 40.0 g of liquid methyl alcohol (CH 3
OH ). Identify the solvent and solute in the
resulting solution.

Answer
solute: HCl(g)

5.1.1 https://chem.libretexts.org/@go/page/423648
solvent: CH 3
OH

LIKE DISSOLVES LIKE


A simple way to predict which compounds will dissolve in other compounds is the phrase "like dissolves like". What this means is that
polar compounds dissolve polar compounds, nonpolar compounds dissolve nonpolar compounds, but polar and nonpolar do not dissolve in
each other.
Even some nonpolar substances dissolve in water but only to a limited degree. Have you ever wondered why fish are able to breathe?
Oxygen gas, a nonpolar molecule, does dissolve in water—it is this oxygen that the fish take in through their gills. The reason we can enjoy
carbonated sodas is also due to a nonpolar compound that dissolves in water. Pepsi-cola and all the other sodas have carbon dioxide gas,
CO , a nonpolar compound, dissolved in a sugar-water solution. In this case, to keep as much gas in solution as possible, the sodas are kept
2

under pressure.
This general trend of "like dissolves like" is summarized in the following table:
Table 5.1.2: Summary of Solubilities
Solute (Polarity of Compound) Solvent (Polarity of Compound) Dominant Intermolecular Force Is Solution Formed?
Polar Polar Dipole-Dipole Force and/or Hydrogen Bond yes
Non-polar Non-polar Dispersion Force yes
Polar Non-polar no
Non-polar Polar no
Ionic Polar Ion-Dipole yes
Ionic Non-polar no

Note that every time charged particles (ionic compounds or polar substances) are mixed, a solution is formed. When particles with no
charges (nonpolar compounds) are mixed, they will form a solution. However, if substances with charges are mixed with other substances
without charges, a solution does not form. When an ionic compound is considered "insoluble", it doesn't necessarily mean the compound is
completely untouched by water. All ionic compounds dissolve to some extent. An insoluble compound just doesn't dissolve in any
noticeable or appreciable amount.
What is it that makes a solute soluble in some solvents but not others?
The answer is intermolecular interactions. The intermolecular interactions include London dispersion forces, dipole-dipole interactions, and
hydrogen bonding (as described in Chapter 10). From experimental studies, it has been determined that if molecules of a solute experience
the same intermolecular forces that the solvent does, the solute will likely dissolve in that solvent. So, NaCl—a very polar substance
because it is composed of ions—dissolves in water, which is very polar, but not in oil, which is generally nonpolar. Nonpolar wax dissolves
in nonpolar hexane, but not in polar water.

Figure 5.1.2: Water (clear liquid) and oil (yellow) do not form liquid solutions. (CC BY-SA 1.0 Generic; Victor Blacus)

 EXAMPLE 5.1.2: POLAR AND NONPOLAR SOLVENTS

Would I be more soluble in CCl or H


2 4 2
O ? Explain your answer.

Solution
I
2
is nonpolar. Of the two solvents, CCl is nonpolar and H
4 2
O is polar, so I would be expected to be more soluble in CCl .
2 4

5.1.2 https://chem.libretexts.org/@go/page/423648
 EXERCISE 5.1.2

Would C 3
H OH
7
be more soluble in CCl or H 4 2
O ? Explain your answer.

Answer
H O
2
, because both experience hydrogen bonding.

 EXAMPLE 5.1.3

Water is considered a polar solvent. Which substances should dissolve in water?


a. methanol (CH OH) 3

b. sodium sulfate (\ce{Na2SO4}\))


c. octane (C H ) 8 18

Solution
Because water is polar, substances that are polar or ionic will dissolve in it.
a. Because of the OH group in methanol, we expect its molecules to be polar. Thus, we expect it to be soluble in water. As both water
and methanol are liquids, the word miscible can be used in place of soluble.
b. Sodium sulfate is an ionic compound, so we expect it to be soluble in water.
c. Like other hydrocarbons, octane is nonpolar, so we expect that it would not be soluble in water.

 EXERCISE 5.1.3: TOLUENE

Toluene (C 6
H5 CH3 ) is widely used in industry as a nonpolar solvent. Which substances should dissolve in toluene?
a. water (H2O)
b. sodium sulfate (Na2SO4)
c. octane (C8H18)

Answer
Octane (C 8
H
18
) will dissolve. It is also non-polar.

SUMMARY
Solutions are composed of a solvent (major component) and a solute (minor component).
“Like dissolves like” is a useful rule for deciding if a solute will be soluble in a solvent.

5.1: Solutions - Homogeneous Mixtures is shared under a CC BY-NC-SA 3.0 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew
& Henry Agnew.
13.2: Solutions - Homogeneous Mixtures by Henry Agnew, Marisa Alviar-Agnew is licensed CC BY-NC-SA 3.0.

5.1.3 https://chem.libretexts.org/@go/page/423648
5.2: Solutions of Solids Dissolved in Water- How to Make Rock Candy
 Learning Objectives
Define electrolytes and non electrolytes
Explain why solutions form.
Discuss the idea of water as the "universal solvent".
Explain how water molecules attract ionic solids when they dissolve in water.

We have learned that solutions can be formed in a variety of combinations using solids, liquids, and gases. We also know that
solutions have constant composition, and that this composition can be varied up to a point to maintain the homogeneous nature of
the solution. But how exactly do solutions form? Why is it that oil and water will not form a solution, and yet vinegar and water
will? Why could we dissolve table salt in water, but not in vegetable oil? The reasons why solutions will form will be explored in
this section, along with a discussion of why water is used most frequently to dissolve substances of various types.

Solubility and Saturation


Table salt (NaCl) readily dissolves in water. In most cases, only a certain maximum amount of solute can be dissolved in a given
amount of solvent. This maximum amount is specified as the solubility of the solute. It is usually expressed in terms of the amount
of solute that can dissolve in 100 g of the solvent at a given temperature. Table 5.2.1 lists the solubilities of some simple ionic
compounds. These solubilities vary widely. NaCl can dissolve up to 31.6 g per 100 g of H2O, while AgCl can dissolve only
0.00019 g per 100 g of H2O.
Table 5.2.1 : Solubilities of Some Ionic Compounds
Solute Solubility (g per 100 g of H2O at 25°C)

AgCl 0.00019

CaCO3 0.0006

KBr 70.7

NaCl 36.1

NaNO3 94.6

When the maximum amount of solute has been dissolved in a given amount of solvent, we say that the solution is saturated with
solute. When less than the maximum amount of solute is dissolved in a given amount of solute, the solution is unsaturated. These
terms are also qualitative terms because each solute has its own solubility. A solution of 0.00019 g of AgCl per 100 g of H2O may
be saturated, but with so little solute dissolved, it is also rather dilute. A solution of 36.1 g of NaCl in 100 g of H2O is also
saturated, but rather concentrated. In some circumstances, it is possible to dissolve more than the maximum amount of a solute in a
solution. Usually, this happens by heating the solvent, dissolving more solute than would normally dissolve at regular temperatures,
and letting the solution cool down slowly and carefully. Such solutions are called supersaturated solutions and are not stable;
given an opportunity (such as dropping a crystal of solute in the solution), the excess solute will precipitate from the solution.The
figure below illustrates the above process and shows the distinction between unsaturated and saturated.

5.2.1 https://chem.libretexts.org/@go/page/423649
Figure 5.2.1 : When 30.0 g of NaCl is added to 100 mL, it all dissolves, forming an unsaturated solution. When 40.0 g is added,
36.0 g dissolves and 4.0 g remains undissolved, forming a saturated solution.

How can you tell if a solution is saturated or unsaturated? If more solute is added and it does not dissolve, then the original solution
was saturated. If the added solute dissolves, then the original solution was unsaturated. A solution that has been allowed to reach
equilibrium, but which has extra undissolved solute at the bottom of the container, must be saturated.

Rock Candy Recipe - Crystallization of S…


S…

Electrolyte Solutions: Dissolved Ionic Solids


When some substances are dissolved in water, they undergo either a physical or a chemical change that yields ions in solution.
These substances constitute an important class of compounds called electrolytes. Substances that do not yield ions when dissolved
are called nonelectrolytes. If the physical or chemical process that generates the ions is essentially 100% efficient (all of the
dissolved compound yields ions), then the substance is known as a strong electrolyte (good conductor). If only a relatively small
fraction of the dissolved substance undergoes the ion-producing process, the substance is a weak electrolyte (does not conduct
electricity as well).
Substances may be identified as strong, weak, or nonelectrolytes by measuring the electrical conductance of an aqueous solution
containing the substance. To conduct electricity, a substance must contain freely mobile, charged species. Most familiar is the
conduction of electricity through metallic wires, in which case the mobile, charged entities are electrons. Solutions may also
conduct electricity if they contain dissolved ions, with conductivity increasing as ion concentration increases. Applying a voltage to
electrodes immersed in a solution permits assessment of the relative concentration of dissolved ions, either quantitatively, by
measuring the electrical current flow, or qualitatively, by observing the brightness of a light bulb included in the circuit (Figure
5.2.1).

5.2.2 https://chem.libretexts.org/@go/page/423649
Figure 5.2.1 : Solutions of nonelectrolytes, such as ethanol, do not contain dissolved ions and cannot conduct electricity. Solutions
of electrolytes contain ions that permit the passage of electricity. The conductivity of an electrolyte solution is related to the
strength of the electrolyte. This diagram shows three separate beakers. Each has a wire plugged into a wall outlet. In each case, the
wire leads from the wall to the beaker and is split resulting in two ends. One end leads to a light bulb and continues on to a
rectangle labeled with a plus sign. The other end leads to a rectangle labeled with a minus sign. The rectangles are in a solution. In
the first beaker, labeled “Ethanol No Conductivity,” four pairs of linked small green spheres suspended in the solution between the
rectangles. In the second beaker, labeled “K C l Strong Conductivity,” six individual green spheres, three labeled plus and three
labeled minus are suspended in the solution. Each of the six spheres has an arrow extending from it pointing to the rectangle
labeled with the opposite sign. In the third beaker, labeled “Acetic acid solution Weak conductivity,” two pairs of joined green
spheres and two individual spheres, one labeled plus and one labeled minus are shown suspended between the two rectangles. The
plus labeled sphere has an arrow pointing to the rectangle labeled minus and the minus labeled sphere has an arrow pointing to the
rectangle labeled plus.

Water and other polar molecules are attracted to ions, as shown in Figure 5.2.2. The electrostatic attraction between an ion and a
molecule with a dipole is called an ion-dipole attraction. These attractions play an important role in the dissolution of ionic
compounds in water.

Figure 5.2.2 : As potassium chloride (KCl) dissolves in water, the ions are hydrated. The polar water molecules are attracted by the
charges on the K+ and Cl− ions. Water molecules in front of and behind the ions are not shown. The diagram shows eight purple
spheres labeled K superscript plus and eight green spheres labeled C l superscript minus mixed and touching near the center of the
diagram. Outside of this cluster of spheres are seventeen clusters of three spheres, which include one red and two white spheres. A
red sphere in one of these clusters is labeled O. A white sphere is labeled H. Two of the green C l superscript minus spheres are
surrounded by three of the red and white clusters, with the red spheres closer to the green spheres than the white spheres. One of
the K superscript plus purple spheres is surrounded by four of the red and white clusters. The white spheres of these clusters are
closest to the purple spheres.
When ionic compounds dissolve in water, the ions in the solid separate and disperse uniformly throughout the solution because
water molecules surround and solvate the ions, reducing the strong electrostatic forces between them. This process represents a
physical change known as dissociation. Under most conditions, ionic compounds will dissociate nearly completely when dissolved,
and so they are classified as strong electrolytes.

 Example 5.2.1: Identifying Ionic Compounds

Which compound(s) will dissolve in solution to separate into ions?


a. LiF

5.2.3 https://chem.libretexts.org/@go/page/423649
b. P 2
F
5

c. C 2
H OH
5

Solution
LiF will separate into ions when dissolved in solution, because it is an ionic compound. P 2
F
5
and C2
H OH
5
are both covalent
and will stay as molecules in a solution.

 Exercise 5.2.1
Which compounds will dissolve in solution to separate into ions?
a. C6H12O11, glucose
b. CCl4
c. CaCl2
d. AgNO3

Answer
c&d

How Temperature Influences Solubility


The solubility of a substance is the amount of that substance that is required to form a saturated solution in a given amount of
solvent at a specified temperature. Solubility is often measured as the grams of solute per 100 g of solvent. The solubility of
sodium chloride in water is 36.0 g per 100 g water at 20 C. The temperature must be specified because solubility varies with
o

temperature. For gases, the pressure must also be specified. Solubility is specific for a particular solvent. We will consider
solubility of material in water as solvent.
The solubility of the majority of solid substances increases as the temperature increases. However, the effect is difficult to predict
and varies widely from one solute to another. The temperature dependence of solubility can be visualized with the help of a
solubility curve, a graph of the solubility vs. temperature (Figure 5.2.4).

Figure 5.2.4 : Solubility curves for several compounds.


Notice how the temperature dependence of NaCl is fairly flat, meaning that an increase in temperature has relatively little effect on
the solubility of NaCl. The curve for KNO , on the other hand, is very steep and so an increase in temperature dramatically
3

increases the solubility of KNO .


3

5.2.4 https://chem.libretexts.org/@go/page/423649
Several substances—HCl, NH , and SO —have solubility that decreases as temperature increases. They are all gases at standard
3 2

pressure. When a solvent with a gas dissolved in it is heated, the kinetic energy of both the solvent and solute increase. As the
kinetic energy of the gaseous solute increases, its molecules have a greater tendency to escape the attraction of the solvent
molecules and return to the gas phase. Therefore, the solubility of a gas decreases as the temperature increases.
Solubility curves can be used to determine if a given solution is saturated or unsaturated. Suppose that 80 g of KNO is added to
3

100 g of water at 30 C. According to the solubility curve, approximately 48 g of KNO will dissolve at 30 C. This means that the
o o
3

solution will be saturated since 48 g is less than 80 g. We can also determine that there will be 80 − 48 = 32 g of undissolved
KNO
3
remaining at the bottom of the container. Now suppose that this saturated solution is heated to 60 C. According to the
o

curve, the solubility of KNO at 60 C is about 107 g. Now the solution is unsaturated since it contains only the original 80 g of
3
o

dissolved solute. Now suppose the solution is cooled all the way down to 0 C. The solubility at 0 C is about 14 g, meaning that
o o

80 − 14 = 66 g of the KNO will re-crystallize.


3

Summary
Solubility is the specific amount of solute that can dissolve in a given amount of solvent.
Saturated and unsaturated solutions are defined.
Ionic compounds dissolve in polar solvents, especially water. This occurs when the positive cation from the ionic solid is
attracted to the negative end of the water molecule (oxygen) and the negative anion of the ionic solid is attracted to the positive
end of the water molecule (hydrogen).
Water is considered the universal solvent since it can dissolve both ionic and polar solutes, as well as some nonpolar solutes (in
very limited amounts).
The solubility of a solid in water increases with an increase in temperature.

Vocabulary
Miscible - Liquids that have the ability to dissolve in each other.
Immiscible - Liquids that do not have the ability to dissolve in each other.
Electrostatic attraction - The attraction of oppositely charged particles.

5.2: Solutions of Solids Dissolved in Water- How to Make Rock Candy is shared under a CK-12 license and was authored, remixed, and/or
curated by Marisa Alviar-Agnew & Henry Agnew.
13.3: Solutions of Solids Dissolved in Water- How to Make Rock Candy by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12.
Original source: https://www.ck12.org/c/chemistry/.

5.2.5 https://chem.libretexts.org/@go/page/423649
5.3: Specifying Solution Concentration- Mass Percent
 Learning Objectives
Express the amount of solute in a solution in various concentration units.

To define a solution precisely, we need to state its concentration: how much solute is dissolved in a certain amount of solvent.
Words such as dilute or concentrated are used to describe solutions that have a little or a lot of dissolved solute, respectively, but
these are relative terms with meanings that depend on various factors.

Introduction
Concentration is the measure of how much of a given substance is mixed with another substance. Solutions are said to be either
dilute or concentrated. When we say that vinegar is 5% acetic acid in water, we are giving the concentration. If we said the mixture
was 10% acetic acid, this would be more concentrated than the vinegar solution.

Figure 5.3.1 : The solution on the left is more concentrated than the solution on the right because there is a greater ratio of solute
(red balls) to solvent (blue balls) particles. The solution particles are closer together. The solution on the right is more dilute (less
concentrated). (CC-SA-BY-3.0 Tracy Poulsen).
A concentrated solution is one in which there is a large amount of solute in a given amount of solvent. A dilute solution is one in
which there is a small amount of solute in a given amount of solvent. A dilute solution is a concentrated solution that has been, in
essence, watered down. Think of the frozen juice containers you buy in the grocery store. To make juice, you have to mix the
frozen juice concentrate from inside these containers with three or four times the container size full of water. Therefore, you are
diluting the concentrated juice. In terms of solute and solvent, the concentrated solution has a lot of solute versus the dilute solution
that would have a smaller amount of solute.
The terms "concentrated" and "dilute" provide qualitative methods of describing concentration. Although qualitative observations
are necessary and have their place in every part of science, including chemistry, we have seen throughout our study of science that
there is a definite need for quantitative measurements in science. This is particularly true in solution chemistry. In this section, we
will explore some quantitative methods of expressing solution concentration.

Mass Percent
There are several ways of expressing the concentration of a solution by using a percentage. The mass/mass percent (% m/m) is
defined as the mass of a solute divided by the mass of a solution times 100:
mass of solute
% m/m = × 100%
mass of solution

mass of solution = mass of solute + mass solvent


If you can measure the masses of the solute and the solution, determining the mass/mass percent is easy. Each mass must be
expressed in the same units to determine the proper concentration.
Suppose that a solution was prepared by dissolving 25.0 g of sugar into 100.0 g of water.

The mass of the solution is


mass of solution = 25.0g sugar + 100.0g water = 125.0 g
The percent by mass would be calculated by:
25.0 g sugar
Percent by mass = × 100% = 20.0% sugar
125.0 g solution

5.3.1 https://chem.libretexts.org/@go/page/423650
 Example 5.3.1

A saline solution with a mass of 355 g has 36.5 g of NaCl dissolved in it. What is the mass/mass percent concentration of the
solution?

Solution
We can substitute the quantities given in the equation for mass/mass percent:
36.5 g
% m/m = × 100% = 10.3%
355 g

 Exercise 5.3.1

A dextrose (also called D-glucose, C6H12O6) solution with a mass of 2.00 × 102 g has 15.8 g of dextrose dissolved in it. What
is the mass/mass percent concentration of the solution?

Answer
7.90 %

Using Mass Percent in Calculations


Sometimes you may want to make up a particular mass of solution of a given percent by mass and need to calculate what mass of
the solute to use. Using mass percent as a conversion can be useful in this type of problem. The mass percent can be expressed as a
g solute 100 gsolution
conversion factor in the form or
100 gsolution g solute

For example, if you need to make 3000.0 g of a 5.00% solution of sodium chloride, the mass of solute needs to be determined.

Solution
Given: 3000.0 g NaCl solution
5.00% NaCl solution
Find: mass of solute = ? g NaCl
Other known quantities: 5.00 g NaCl is to 100 g solution
The appropriate conversion factor (based on the given mass percent) can be used follows:

To solve for the mass of NaCl, the given mass of solution is multiplied by the conversion factor.
5.00 g N aC l
gN aC l = 3, 000.0 g N aC l solution × = 150.0g N aC l
100 g N aC l solution

You would need to weigh out 150 g of NaCl and add it to 2850 g of water. Notice that it was necessary to subtract the mass of the
NaCl (150 g) from the mass of solution (3000 g) to calculate the mass of the water that would need to be added.

5.3.2 https://chem.libretexts.org/@go/page/423650
 Exercise 5.3.1

What is the amount (in g) of hydrogen peroxide (H2O2) needed to make a 6.00 kg, 3.00 % (by mass) H2O2 solution?

Answer
180 g H2O2

Contributors and Attributions

5.3: Specifying Solution Concentration- Mass Percent is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa
Alviar-Agnew & Henry Agnew.
13.5: Specifying Solution Concentration- Mass Percent by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

5.3.3 https://chem.libretexts.org/@go/page/423650
5.4: Specifying Solution Concentration- Molarity
 Learning Objectives
Use molarity to determine quantities in chemical reactions.
Use molarity as a conversion factor in calculations.

Another way of expressing concentration is to give the number of moles of solute per unit volume of solution. Of all the quantitative measures
of concentration, molarity is the one used most frequently by chemists. Molarity is defined as the number of moles of solute per liter of
solution.
number of moles of solute
molarity = (5.4.1)
number of liters of solution

The symbol for molarity is M or moles/liter. Chemists also use square brackets to indicate a reference to the molarity of a substance. For
example, the expression [ Ag ] refers to the molarity of the silver ion in solution. Solution concentrations expressed in molarity are the easiest
+

to perform calculations with, but the most difficult to make in the lab. Such concentration units are useful for discussing chemical reactions in
which a solute is a product or a reactant. Molar mass can then be used as a conversion factor to convert amounts in moles to amounts in grams.
It is important to remember that “mol” in this expression refers to moles of solute and that “L” refers to liters of solution. For example, if you
have 1.5 mol of NaCl dissolved in 0.500 L of solution, its molarity is
1.5 mol NaCl
= 3.0 M NaCl
0.500 L solution

Sometimes (aq) is added when the solvent is water, as in “3.0 M NaCl (aq).” This is read as "a 3.00 molar sodium chloride solution," meaning
that there are 3.00 moles of NaOH dissolved per one liter of solution.

Be sure to note that molarity is calculated as the total volume of the entire solution, not just volume
of solvent! The solute contributes to total volume.
If the quantity of the solute is given in mass units, you must convert mass units to mole units before using the definition of molarity to
calculate concentration. For example, what is the molar concentration of a solution of 22.4 g of HCl dissolved in 1.56 L?
Step 1: First, convert the mass of solute to moles using the molar mass of HCl (36.5 g/mol):
1 mol H C l
22.4 gH C l × = 0.614 mol H C l
36.5 gH C l

Step 2: Now we can use the definition of molarity to determine a concentration:


0.614 mol H C l
M = = 0.394 M H C l
1.56L solution

Before a molarity concentration can be calculated, the amount of the solute must be expressed in moles, and the volume of the solution must
be expressed in liters, as demonstrated in the following example.

 Example 5.4.1

A solution is prepared by dissolving 42.23 g of NH 4


Cl into enough water to make 500.0 mL of solution. Calculate its molarity.

Solution
Solutions to Example 13.6.1
Steps for Problem Solving

Given:
Identify the "given" information and what the problem is asking you to Mass = 42.23 g NH Cl 4

"find." Volume solution = 500.0 mL = 0.5000 L


Find: Molarity = ? M

List other known quantities. Molar mass NH 4


Cl = 53.50 g/mol

5.4.1 https://chem.libretexts.org/@go/page/423651
Steps for Problem Solving

1. The mass of the ammonium chloride is first converted to moles.

Plan the problem.

2. Then the molarity is calculated by dividing by liters. Note the given


volume has been converted to liters.
mol N H4 Cl
M =
L solution

Now substitute the known quantities into the equation and solve.
1 mol NH Cl
4
42.23 g NH Cl × = 0.7893 mol NH Cl
4 4
Cancel units and calculate. 53.50 g NH Cl
4

0.7893 mol NH Cl
4
= 1.579 M
0.5000 L solution

The molarity is 1.579 M, meaning that a liter of the solution would contain
Think about your result.
1.579 mol NH Cl. Four significant figures are appropriate.
4

 Exercise 5.4.1A

What is the molarity of a solution made when 66.2 g of C6H12O6 are dissolved to make 235 mL of solution?

Answer
1.57 M C6H12O6

 Exercise 5.4.1B

What is the concentration, in mol/L, where 137 g of NaCl has been dissolved in enough water to make 500 mL of solution?

Answer
4.69 M NaCl

Using Molarity in Calculations


Concentration can be a conversion factor between the amount of solute and the amount of solution or solvent (depending on the definition of
the concentration unit). As such, concentrations can be useful in a variety of stoichiometry problems. In many cases, it is best to use the
original definition of the concentration unit; it is that definition that provides the conversion factor.
A simple example of using a concentration unit as a conversion factor is one in which we use the definition of the concentration unit and
rearrange; we can do the calculation again as a unit conversion, rather than as a definition.

 Example 5.4.2: Determining Moles of Solute, Given the Concentration and Volume of a Solution

For example, suppose we ask how many moles of solute are present in 0.108 L of a 0.887 M NaCl solution. Because 0.887 M means 0.887
mol/L, we can use this second expression for the concentration as a conversion factor:

Solution

5.4.2 https://chem.libretexts.org/@go/page/423651
0.887 mol N aC l
0.108 L solution × = 0.0958 mol N aC l
1L solution

If we used the definition approach, we get the same answer, but now we are using conversion factor skills.

Like any other conversion factor that relates two different types of units, the reciprocal of the concentration can be also used as a conversion
factor.

 Example 5.4.3: Determining Volume of a Solution, Given the Concentration and Moles of Solute

Using concentration as a conversion factor, how many liters of 2.35 M CuSO4 are needed to obtain 4.88 mol of CuSO4?

Solution
This is a one-step conversion, but the concentration must be written as the reciprocal for the units to work out:

1 L solution
4.88 mol C uSO4 × = 2.08 L of solution
2.35 mol C uSO4

In a laboratory situation, a chemist must frequently prepare a given volume of solutions of a known molarity. The task is to calculate the mass
of the solute that is necessary. The molarity equation can be rearranged to solve for moles, which can then be converted to grams. The
following example illustrates this.

 Example 5.4.4

A chemist needs to prepare 3.00 L of a 0.250 M solution of potassium permanganate (KMnO ). What mass of KMnO does she need to
4 4

make the solution?

Solution
Solutions to Example 13.6.4
Steps for Problem Solving

Given:
Identify the "given" information and what the problem is asking you to Molarity = 0.250 M
"find." Volume = 3.00 L
Find: Mass KMnO =? g
4

Molar mass KMnO = 158.04 g/mol


List other known quantities. 4

0.250 mol KMnO4 to 1 L of KMnO4 solution

Plan the problem.

Now substitute the known quantities into the equation and solve.
mol KMnO = 0.250 M KMnO × 3.00 L = 0.750 mol KM
4 4

Cancel units and calculate. 0.250 mol KMnO


4 158.04 g KMnO
4
3.00 L solution × × = 119 g KMnO
1 L solution 1 mol KMnO
4

When 119 g of potassium permanganate is dissolved into water to make


Think about your result.
3.00 L of solution, the molarity is 0.250 M.

5.4.3 https://chem.libretexts.org/@go/page/423651
 Exercise 5.4.4A
Using concentration as a conversion factor, how many liters of 0.0444 M CH2O are needed to obtain 0.0773 mol of CH2O?

Answer
1.74 L

 Exercise 5.4.4B
Answer the problems below using concentration as a conversion factor.
a. What mass of solute is present in 1.08 L of 0.0578 M H2SO4?
b. What volume of 1.50 M HCl solution contains 10.0 g of hydrogen chloride?

Answer a
6.12 g
Answer b
183 mL or 0.183L

 How to Indicate Concentration


Square brackets are often used to represent concentration, e.g., [NaOH] = 0.50 M.
Use the capital letter M for molarity, not a lower case m (this is a different concentration unit called molality).

How To Prepare Solutions

Watch as the Flinn Scientific Tech Staff demonstrates "How To Prepare Solutions."
It is important to note that there are many different ways you can set up and solve your chemistry equations. Some students prefer to answer
multi-step calculations in one long step, while others prefer to work out each step individually. Neither method is necessarily better or worse
than the other method—whichever makes the most sense to you is the one that you should use. In this text, we will typically use unit analysis
(also called dimension analysis or factor analysis).

Contributors and Attributions


Peggy Lawson (Oxbow Prairie Heights School). Funded by Saskatchewan Educational Technology Consortium.

5.4: Specifying Solution Concentration- Molarity is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew &
Henry Agnew.
13.6: Specifying Solution Concentration- Molarity by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

5.4.4 https://chem.libretexts.org/@go/page/423651
5.5: SOLUTION DILUTION
 LEARNING OBJECTIVES
Explain how concentrations can be changed in the lab.
Understand how stock solutions are used in the laboratory.

We are often concerned with how much solute is dissolved in a given amount of solution. We will begin our discussion of solution
concentration with two related and relative terms: dilute and concentrated.
A dilute solution is one in which there is a relatively small amount of solute dissolved in the solution.
A concentrated solution contains a relatively large amount of solute.
These two terms do not provide any quantitative information (actual numbers), but they are often useful in comparing solutions in a more
general sense. These terms also do not tell us whether or not the solution is saturated or unsaturated, or whether the solution is "strong" or
"weak". These last two terms will have special meanings when we discuss acids and bases, so be careful not to confuse them.

STOCK SOLUTIONS
It is often necessary to have a solution with a concentration that is very precisely known. Solutions containing a precise mass of solute in a
precise volume of solution are called stock (or standard) solutions. To prepare a standard solution, a piece of lab equipment called a
volumetric flask should be used. These flasks range in size from 10 mL to 2000 mL and are carefully calibrated to a single volume. On the
narrow stem is a calibration mark. The precise mass of solute is dissolved in a bit of the solvent, and this is added to the flask. Then,
enough solvent is added to the flask until the level reaches the calibration mark.
Often, it is convenient to prepare a series of solutions of known concentrations by first preparing a single stock solution, as described in the
previous section. Aliquots (carefully measured volumes) of the stock solution can then be diluted to any desired volume. In other cases, it
may be inconvenient to weigh a small mass of sample accurately enough to prepare a small volume of a dilute solution. Each of these
situations requires that a solution be diluted to obtain the desired concentration.

DILUTIONS OF STOCK (OR STANDARD) SOLUTIONS


Imagine we have a salt water solution with a certain concentration. That means we have a certain amount of salt (a certain mass or a certain
number of moles) dissolved in a certain volume of solution. Next, we will dilute this solution. This is done by adding more water, not more
salt:

Before Dilution and After Dilution


The molarity of solution 1 is
moles 1
M1 =
liter 1

and the molarity of solution 2 is


moles 2
M2 =
liter 2

rearrange the equations to find moles:

moles 1 = M1 liter 1

and

moles 2 = M2 liter 2

What stayed the same and what changed between the two solutions? By adding more water, we changed the volume of the solution. Doing
so also changed its concentration. However, the number of moles of solute did not change. So,

moles1 = moles2

5.5.1 https://chem.libretexts.org/@go/page/423652
Therefore

M1 V1 = M2 V2 (5.5.1)

where
M1 and M are the concentrations of the original and diluted solutions
2

V1 and V are the volumes of the two solutions


2

Preparing dilutions is a common activity in the chemistry lab and elsewhere. Once you understand the above relationship, the calculations
are simple.
Suppose that you have 100. mL of a 2.0 M solution of HCl . You dilute the solution by adding enough water to make the solution volume
500. mL . The new molarity can easily be calculated by using the above equation and solving for M . 2

M1 × V1 2.0 M × 100. mL
M2 = = = 0.40 M HCl
V2 500. mL

The solution has been diluted by one-fifth since the new volume is five times as great as the original volume. Consequently, the molarity is
one-fifth of its original value.
Another common dilution problem involves calculating what amount of a highly concentrated solution is required to make a desired
quantity of solution of lesser concentration. The highly concentrated solution is typically referred to as the stock solution.

 EXAMPLE 5.5.1: DILUTING NITRIC ACID

Nitric acid (HNO ) is a powerful and corrosive acid. When ordered from a chemical supply company, its molarity is 16 M . How much
3

of the stock solution of nitric acid needs to be used to make 8.00 L of a 0.50 M solution?

Solution
Solutions to Example13.7.1
Steps for
Problem
Solving
Identify the
"given"
Given:
information
M1, Stock HNO = 16 M
and what
3

V = 8.00 L
the
2

M = 0.50 M
problem is
2

Find: Volume stock HNO (V1 ) =? L


asking you 3

to "find."
List other
known none
quantities.
First, rearrange the equation algebraically to solve for V .1

Plan the
M2 × V2
problem. V1 =
M1

Now substitute the known quantities into the equation and solve.
Calculate
and cancel 0.50 M × 8.00 L
V1 = = 0.25 L (5.5.2)
units. 16 M

Think
0.25 L (250 mL) of the stock HNO needs to be diluted with water to a final volume of 8.00 L. The dilution is by a factor of 32 to go from 16 M to
about your 3

0.5 M.
result.

 EXERCISE 5.5.1

A 0.885 M solution of KBr with an initial volume of 76.5 mL has more water added until its concentration is 0.500 M. What is the new
volume of the solution?

5.5.2 https://chem.libretexts.org/@go/page/423652
Answer
135.4 mL
Note that the calculated volume will have the same dimensions as the input volume, and dimensional analysis tells us that in this case
we don't need to convert to liters, since L cancels when we divide M (mol/L) by M (mol/L).

DILUTING AND MIXING SOLUTIONS


Diluting and Mixing Solutions

How to Dilute a Sol…


Sol…

How to Dilute a Solution by CarolinaBiological

This page titled 5.5: Solution Dilution is shared under a CC BY-SA 4.0 license and was authored, remixed, and/or curated by Ed Vitz, John W. Moore,
Justin Shorb, Xavier Prat-Resina, Tim Wendorff, & Adam Hahn.
13.7: Solution Dilution by Ed Vitz, John W. Moore, Justin Shorb, Xavier Prat-Resina, Tim Wendorff, & Adam Hahn is licensed CC BY-SA 4.0.

5.5.3 https://chem.libretexts.org/@go/page/423652
5.6: SOLUTION STOICHIOMETRY
 LEARNING OBJECTIVES

Determine amounts of reactants or products in aqueous solutions.

As we learned previously, double replacement reactions involve the reaction between ionic compounds in solution and, in the course of the reaction, the ions in
the two reacting compounds are “switched” (they replace each other). Because these reactions occur in aqueous solution, we can use the concept of molarity to
directly calculate the number of moles of reactants or products that will be formed, and therefore their amounts (i.e. volume of solutions or mass of
precipitates).

As an example, lead (II) nitrate and sodium chloride react to form sodium nitrate and the insoluble compound, lead (II) chloride.
Pb(NO ) (aq) + 2 NaCl(aq) → PbCl (s) + 2 NaNO (aq) (5.6.1)
3 2 2 3

In the reaction shown above, if we mixed 0.123 L of a 1.00 M solution of NaCl with 1.50 M solution of Pb(NO )
3 2
, we could calculate the volume of
Pb(NO ) solution needed to completely precipitate the Pb ions. 2+

3 2

The molar concentration can also be expressed as the following:


1.00 mol NaCl
1.00 M NaCl =
1 L NaCl solution

and
1.50 mol Pb(NO )
3 2
1.50 M Pb(NO ) =
3 2
1 L Pb(NO ) solution
3 2

First, we must examine the reaction stoichiometry in the balanced reaction (Equation 5.6.1 ). In this reaction, one mole of Pb(NO 3
)
2
reacts with two moles of
NaCl to give one mole of PbCl precipitate. Thus, the concept map utilizing the stoichiometric ratios is:
2

so the volume of lead (II) nitrate that reacted is calculated as:

1.00 mol NaCl 1 mol Pb(NO ) 1 L Pb(NO ) solution


3 2 3 2
0.123 L NaCl solution × × × = 0.041 Pb(NO )
3 2
1 L NaCl solution 2 mol NaCl 1.5 mol Pb(NO )
3 2

L solution

This volume makes intuitive sense for two reasons: (1) the number of moles of Pb(NO ) required is half of the number of moles of NaCl , based off of the
3 2

stoichiometry in the balanced reaction (Equation 5.6.1 ); (2) the concentration of Pb(NO ) solution is 50% greater than the NaCl solution, so less volume is
3 2

needed.

 EXAMPLE 5.6.1

What volume (in L) of 0.500 M sodium sulfate will react with 275 mL of 0.250 M barium chloride to completely precipitate all Ba 2+
in the solution?

Solution

5.6.1 https://chem.libretexts.org/@go/page/423653
Solutions to Example 13.8.1
Steps for
Problem Example 5.6.1
Solving
Identify the
"given" Given: 275 mL BaCl2
information 0.250 molBaC l2
0.250 M BaCl or
and what 2
1 L BaC l2 solution

the 0.500 M Na SO or
0.500 molN a2 S O4

2 4
problem is 1 L N a2 S O4 solution

asking you Find: Volume Na 2


SO
4
solution.
to "find."
Set up and
balance the Na SO (aq) + BaCl (aq) ⟶ BaSO (s) + 2
2 4 2 –
NaCl(aq)
4

chemical An insoluble product is formed after the reaction.


equation.
List other
1 mol of Na2SO4 to 1 mol BaCl2
known
1000 mL = 1 L
quantities.
Prepare a
concept
map and
use the
proper
conversion
factor.

Cancel 1 L 0.250 mol BaC l2 1 mol N a2 S O4 1 L N a2 S O4 solution


275 mL BaC l2 solution × × × ×
units and 1000 mL 1 L BaC l2 solution 1 mol BaC l2 0.500 molN a2 S O4

calculate. = 0.1375 L sodium sulfate


Think
The lesser amount (almost half) of sodium sulfate is to be expected as it is more concentrated than barium chloride.
about your
Also, the units are correct.
result.

 EXERCISE 5.6.1

What volume of 0.250 M lithium hydroxide will completely react with 0.500 L of 0.250 M of sulfuric acid solution?

Answer
0.250 L LiOH solution

This page titled 5.6: Solution Stoichiometry is shared under a CC BY-SA 4.0 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew, Henry Agnew, Paul
Young, & Paul Young (ChemistryOnline.com) .
13.8: Solution Stoichiometry by Henry Agnew, Marisa Alviar-Agnew, Paul Young is licensed CC BY-SA 4.0.

5.6.2 https://chem.libretexts.org/@go/page/423653
5.7: Electrolytes
Skills to Develop
Define and give examples of electrolytes
Distinguish between the physical and chemical changes that accompany dissolution of ionic and covalent electrolytes
Relate electrolyte strength to solute-solvent attractive forces

When some substances are dissolved in water, they undergo either a physical or a chemical change that yields ions in solution.
These substances constitute an important class of compounds called electrolytes. Substances that do not yield ions when dissolved
are called nonelectrolytes. If the physical or chemical process that generates the ions is essentially 100% efficient (all of the
dissolved compound yields ions), then the substance is known as a strong electrolyte. If only a relatively small fraction of the
dissolved substance undergoes the ion-producing process, it is called a weak electrolyte.
Substances may be identified as strong, weak, or nonelectrolytes by measuring the electrical conductance of an aqueous solution
containing the substance. To conduct electricity, a substance must contain freely mobile, charged species. Most familiar is the
conduction of electricity through metallic wires, in which case the mobile, charged entities are electrons. Solutions may also
conduct electricity if they contain dissolved ions, with conductivity increasing as ion concentration increases. Applying a voltage to
electrodes immersed in a solution permits assessment of the relative concentration of dissolved ions, either quantitatively, by
measuring the electrical current flow, or qualitatively, by observing the brightness of a light bulb included in the circuit (Figure
5.7.1).

Figure 5.7.1 : Solutions of nonelectrolytes such as ethanol do not contain dissolved ions and cannot conduct electricity. Solutions
of electrolytes contain ions that permit the passage of electricity. The conductivity of an electrolyte solution is related to the
strength of the electrolyte.

Ionic Electrolytes
Water and other polar molecules are attracted to ions, as shown in Figure 5.7.2. The electrostatic attraction between an ion and a
molecule with a dipole is called an ion-dipole attraction. These attractions play an important role in the dissolution of ionic
compounds in water.

5.7.1 https://chem.libretexts.org/@go/page/423654
Figure 5.7.2 : As potassium chloride (KCl) dissolves in water, the ions are hydrated. The polar water molecules are attracted by
the charges on the K+ and Cl− ions. Water molecules in front of and behind the ions are not shown.
When ionic compounds dissolve in water, the ions in the solid separate and disperse uniformly throughout the solution because
water molecules surround and solvate the ions, reducing the strong electrostatic forces between them. This process represents a
physical change known as dissociation. Under most conditions, ionic compounds will dissociate nearly completely when dissolved,
and so they are classified as strong electrolytes.
Let us consider what happens at the microscopic level when we add solid KCl to water. Ion-dipole forces attract the positive
(hydrogen) end of the polar water molecules to the negative chloride ions at the surface of the solid, and they attract the negative
(oxygen) ends to the positive potassium ions. The water molecules penetrate between individual K+ and Cl− ions and surround
them, reducing the strong interionic forces that bind the ions together and letting them move off into solution as solvated ions, as
Figure shows. The reduction of the electrostatic attraction permits the independent motion of each hydrated ion in a dilute solution,
resulting in an increase in the disorder of the system as the ions change from their fixed and ordered positions in the crystal to
mobile and much more disordered states in solution. This increased disorder is responsible for the dissolution of many ionic
compounds, including KCl, which dissolve with absorption of heat.
In other cases, the electrostatic attractions between the ions in a crystal are so large, or the ion-dipole attractive forces between the
ions and water molecules are so weak, that the increase in disorder cannot compensate for the energy required to separate the ions,
and the crystal is insoluble. Such is the case for compounds such as calcium carbonate (limestone), calcium phosphate (the
inorganic component of bone), and iron oxide (rust).

Covalent Electrolytes
Pure water is an extremely poor conductor of electricity because it is only very slightly ionized—only about two out of every 1
billion molecules ionize at 25 °C. Water ionizes when one molecule of water gives up a proton to another molecule of water,
yielding hydronium and hydroxide ions.
+ −
H2 O(l) + H2 O(l) ⇌ H3 O (aq)
+ OH (aq)
(5.7.1)

In some cases, we find that solutions prepared from covalent compounds conduct electricity because the solute molecules react
chemically with the solvent to produce ions. For example, pure hydrogen chloride is a gas consisting of covalent HCl molecules.
This gas contains no ions. However, when we dissolve hydrogen chloride in water, we find that the solution is a very good
conductor. The water molecules play an essential part in forming ions: Solutions of hydrogen chloride in many other solvents, such
as benzene, do not conduct electricity and do not contain ions.
Hydrogen chloride is an acid, and so its molecules react with water, transferring H+ ions to form hydronium ions (H 3O
+
) and
chloride ions (Cl−):

This reaction is essentially 100% complete for HCl (i.e., it is a strong acid and, consequently, a strong electrolyte). Likewise, weak
acids and bases that only react partially generate relatively low concentrations of ions when dissolved in water and are classified as
weak electrolytes. The reader may wish to review the discussion of strong and weak acids provided in the earlier chapter of this
text on reaction classes and stoichiometry.

Summary
Substances that dissolve in water to yield ions are called electrolytes. Electrolytes may be covalent compounds that chemically
react with water to produce ions (for example, acids and bases), or they may be ionic compounds that dissociate to yield their
constituent cations and anions, when dissolved. Dissolution of an ionic compound is facilitated by ion-dipole attractions between
the ions of the compound and the polar water molecules. Soluble ionic substances and strong acids ionize completely and are
strong electrolytes, while weak acids and bases ionize to only a small extent and are weak electrolytes. Nonelectrolytes are
substances that do not produce ions when dissolved in water.

5.7.2 https://chem.libretexts.org/@go/page/423654
Glossary
dissociation
physical process accompanying the dissolution of an ionic compound in which the compound’s constituent ions are solvated
and dispersed throughout the solution

electrolyte
substance that produces ions when dissolved in water

ion-dipole attraction
electrostatic attraction between an ion and a polar molecule

nonelectrolyte
substance that does not produce ions when dissolved in water

strong electrolyte
substance that dissociates or ionizes completely when dissolved in water

weak electrolyte
substance that ionizes only partially when dissolved in water

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-
2bd...a7ac8df6@9.110).

5.7: Electrolytes is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.

5.7.3 https://chem.libretexts.org/@go/page/423654
5.8: Acids- Properties and Examples
 Learning Objectives
Examine properties of acids.

Many people enjoy drinking coffee. A cup first thing in the morning helps start the day. But keeping the coffee maker clean can be
a problem. Lime deposits build up after a while and slow down the brewing process. The best cure for this is to put vinegar (dilute
acetic acid) in the pot and run it through the brewing cycle. The vinegar dissolves the deposits and cleans the maker, which will
speed up the brewing process back to its original rate. Just be sure to run water through the brewing process after the vinegar, or
you will get some really horrible coffee.

Acids
Acids are very common in some of the foods that we eat. Citrus fruits such as oranges and lemons contain citric acid and ascorbic
acid, which is better known as vitamin C. Carbonated sodas contain phosphoric acid. Vinegar contains acetic acid. Your own
stomach utilizes hydrochloric acid to digest food. Acids are a distinct class of compounds because of the properties of their aqueous
solutions as outlined below:
1. Aqueous solutions of acids are electrolytes, meaning that they conduct electrical current. Some acids are strong electrolytes
because they ionize completely in water, yielding a great many ions. Other acids are weak electrolytes that exist primarily in a
non-ionized form when dissolved in water.
2. Acids have a sour taste. Lemons, vinegar, and sour candies all contain acids.
3. Acids change the color of certain acid-base indicates. Two common indicators are litmus and phenolphthalein. Blue litmus turns
red in the presence of an acid, while phenolphthalein turns colorless.
4. Acids react with active metals to yield hydrogen gas. Recall that an activity series is a list of metals in descending order of
reactivity. Metals that are above hydrogen in the activity series will replace the hydrogen from an acid in a single-replacement
reaction, as shown below:
Zn (s) + H SO (aq) → ZnSO (aq) + H (g) (5.8.1)
2 4 4 2

5. Acids react with bases to produce a salt compound and water. When equal moles of an acid and a base are combined, the acid is
neutralized by the base. The products of this reaction are an ionic compound, which is labeled as a salt, and water.
It should not be hard for you to name several common acids (but you might find that listing bases is a little more difficult). Below
is a partial list of some common acids, along with some chemical formulas:
Table 5.8.1 : Common Acids and Their Uses
Chemist Name Common Name Uses

Used in cleaning (refining) metals, in


muriatic acid (used in pools) and stomach acid
hydrochloric acid, HCl maintenance of swimming pools, and for
is HCl
household cleaning.
Used in car batteries, and in the manufacture of
sulfuric acid, H2SO4
fertilizers.
Used in the manufacture of fertilizers,
nitric acid, HNO3
explosives and in extraction of gold.

acetic acid, HC2H3O2 vinegar Main ingredient in vinegar.

carbonic acid, H2CO3 responsible for the "fizz" in carbonated drinks As an ingredient in carbonated drinks.

Used in food and dietary supplements. Also


citric acid, C6H8O7 added as an acidulant in creams, gels, liquids,
and lotions.

acetylsalicylic acid, C6H4(OCOCH3)CO2H aspirin The active ingredient in aspirin.

What exactly makes an acid an acid, and what makes a base act as a base? Take a look at the formulas given in the above table and
take a guess.

5.8.1 https://chem.libretexts.org/@go/page/423655
Hydrochloric Acid
Hydrochloric acid is a corrosive, strong mineral acid with many industrial uses. A colorless, highly pungent solution of hydrogen
chloride (HCl) in water. Hydrochloric acid is usually prepared by treating HCl with water.
+ −
HCl(g) + H O(l) ⟶ H O (aq) + Cl (aq)
2 3

Hydrochloric acid can therefore be used to prepare chloride salts. Hydrochloric acid is a strong acid, since it is completely
dissociated in water. Hydrochloric acid is the preferred acid in titration for determining the amount of bases.

Sulfuric Acid
Sulfuric acid is a highly corrosive strong mineral acid with the molecular formula H SO . Sulfuric acid is a diprotic acid and has a
2 4

wide range of applications including use in domestic acidic drain cleaners,[as an electrolyte in lead-acid batteries, and in various
cleaning agents. It is also a central substance in the chemical industry.

Figure 5.8.1 : Drops of concentrated sulfuric acid rapidly decompose a piece of cotton towel by dehydration. (CC BY-SA 3.0; Toxic
Walker).
Because the hydration of sulfuric acid is thermodynamically favorable (and is highly exothermic) and the affinity of it for water is
sufficiently strong, sulfuric acid is an excellent dehydrating agent. Concentrated sulfuric acid has a very powerful dehydrating
property, removing water (H O ) from other compounds including sugar and other carbohydrates and producing carbon, heat,
2

steam. Sulfuric acid behaves as a typical acid in its reaction with most metals by generating hydrogen gas (Equation 5.8.2).
M + H SO → M(SO ) + H (5.8.2)
2 4 4 2

Nitric Acid
Nitric acid (HNO ) is a highly corrosive mineral acid and is also commonly used as a strong oxidizing agent. Nitric acid is
3

normally considered to be a strong acid at ambient temperatures. Nitric acid can be made by reacting nitrogen dioxide (NO (g)) 2

with water.

3 NO (g) + H O(l) → 2 HNO (ag) + NO(g)


2 2 3

Nitric acid reacts with most metals, but the details depend on the concentration of the acid and the nature of the metal. Dilute nitric
acid behaves as a typical acid in its reaction with most metals (e.g., nitric acid with magnesium, manganese or zinc will liberate H 2

gas):

Mg + 2 HNO → Mg (NO ) +H
3 3 2 2

Mn + 2 HNO → Mn(NO ) +H
3 3 2 2

Zn + 2 HNO → Zn(NO ) +H
3 3 2 2

Nitric acid is a corrosive acid and a powerful oxidizing agent. The major hazard it poses is chemical burn, as it carries out acid
hydrolysis with proteins (amide) and fats (ester) which consequently decomposes living tissue (Figure 5.8.2). Concentrated nitric
acid stains human skin yellow due to its reaction with the keratin

5.8.2 https://chem.libretexts.org/@go/page/423655
Figure 5.8.2 : Second degree burn caused by nitric acid. (CC BY-SA 3.0; Alcamán).

Carbonic Acid
Carbonic acid is a chemical compound with the chemical formula H CO and is also a name sometimes given to solutions of
2 3

carbon dioxide in water (carbonated water), because such solutions contain small amounts of H CO (aq). Carbonic acid, which is
2 3

a weak acid, forms two kinds of salts: the carbonates and the bicarbonates. In geology, carbonic acid causes limestone to dissolve,
producing calcium bicarbonate—which leads to many limestone features such as stalactites and stalagmites. Carbonic acid is a
polyprotic acid, specifically it is diprotic, meaning that it has two protons which may dissociate from the parent molecule.
When carbon dioxide dissolves in water, it exists in chemical equilibrium (discussed in Chapter 15), producing carbonic acid:

CO +H O −
↽⇀
− H CO
2 2 2 3

The reaction can be pushed to favor the reactants to generate CO (g)


2
from solution, which is key to the bubbles observed in
carbonated beverages (Figure 5.8.3).
Figure 5.8.3 : A glass of sparkling water. (CC BY-SA 3.0; Nevit Dilmen).

Formic Acid
Formic acid (HCO H ) is the simplest carboxylic acid and is an important intermediate in chemical synthesis and occurs naturally,
2

most notably in some ants. The word "formic" comes from the Latin word for ant, formica, referring to its early isolation by the
distillation of ant bodies. Formic acid occurs widely in nature as its conjugate base formate.

Citric Acid
Citric acid (C H O ) is a weak organic tricarboxylic acid that occurs naturally in citrus fruits. The citrate ion is an intermediate in
6 8 7

the TCA cycle (Krebs cycle), a central metabolic pathway for animals, plants and bacteria. Because it is one of the stronger edible
acids, the dominant use of citric acid is used as a flavoring and preservative in food and beverages, especially soft drinks.

Figure 5.8.4 : Lemons, oranges, limes, and other citrus fruits possess high concentrations of citric acid (CC BY-SA 2.5; André
Karwath).

Acetylsalicylic Acid
Acetylsalicylic acid (also known as aspirin) is a medication used to treat pain, fever, and inflammation. Aspirin, in the form of
leaves from the willow tree, has been used for its health effects for at least 2,400 years.

5.8.3 https://chem.libretexts.org/@go/page/423655
Figure 5.8.5 : Ball-and-stick model of the aspirin molecule. (Public Domain; Ben Mills).
Aspirin is a white, crystalline, weakly acidic substance.

Summary
A brief summary of key aspects of several acids commonly encountered by students was given. Acids are a distinct class of
compounds because of the properties of their aqueous solutions.

Contributions & Attributions

Peggy Lawson (Oxbow Prairie Heights School). Funded by Saskatchewan Educational Technology Consortium.

5.8: Acids- Properties and Examples is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew &
Henry Agnew.
14.2: Acids- Properties and Examples by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

5.8.4 https://chem.libretexts.org/@go/page/423655
5.9: Bases- Properties and Examples
 Learning Objectives
Examine properties of bases.

Perhaps you have eaten too much pizza and felt very uncomfortable hours later. This feeling is due to excess stomach acid being
produced. The discomfort can be dealt with by taking an antacid. The base in the antacid will react with the HCl in the stomach
and neutralize it, taking care of that unpleasant feeling.

Bases
Bases have properties that mostly contrast with those of acids.
1. Aqueous solutions of bases are also electrolytes. Bases can be either strong or weak, just as acids can.
2. Bases often have a bitter taste and are found in foods less frequently than acids. Many bases, like soaps, are slippery to the
touch.
3. Bases also change the color of indicators. Litmus turns blue in the presence of a base, while phenolphthalein turns pink.
4. Bases do not react with metals in the way that acids do.
5. Bases react with acids to produce a salt and water.

Figure 5.9.1 : Phenolphthalein indicator in presence of base.

 Warning!

Tasting chemicals and touching them are NOT good lab practices and should be avoided—in other words—don't do this at
home.

Bases are less common as foods, but they are nonetheless present in many household products. Many cleaners contain ammonia, a
base. Sodium hydroxide is found in drain cleaner. Antacids, which combat excess stomach acid, are comprised of bases such as
magnesium hydroxide or sodium hydrogen carbonate. Various common bases and corresponding uses are given in Table 5.9.2.
Table 5.9.1 : Common Bases and Corresponding Uses
Some Common Bases Uses

sodium hydroxide, NaOH Used in the manufacture of soaps and detergents, and as the main
(lye or caustic soda) ingredient in oven and drain cleaners.

potassium hydroxide, KOH Used in the production of liquid soaps and soft soaps. Used in alkaline
(lye or caustic potash) batteries.

magnesium hydroxide, Mg(OH)2 Used as an ingredient in laxatives, antacids, and deodorants. Also used
(milk of magnesia) in the neutralization of acidic wastewater.

calcium hydroxide, Ca(OH)2 Used in the manufacture of cement and lime water. Also, added to
(slaked lime) neutralize acidic soil.

aluminum hydroxide Used in water purification and as an ingredient in antacids.

Used as a building block for the synthesis of many pharmaceutical


ammonia, NH3 products and in many commercial cleaning products. Used in the
manufacture of fertilizers.

5.9.1 https://chem.libretexts.org/@go/page/423656
Sodium Hydroxide
Sodium hydroxide, also known as lye and caustic soda, is an inorganic compound with formula NaOH . It is a white solid ionic
compound consisting of sodium cations Na and hydroxide anions OH .
+ −

Dissolution of solid sodium hydroxide in water is a highly exothermic reaction:


+ −
NaOH(s) → Na (aq) + OH (aq)

The resulting solution is usually colorless and odorless and feels slippery when it comes in contact with skin.

Figure 5.9.1 : Sample of sodium hydroxide as pellets in a watch glass. (Public Domain; Walkerma.)

Potassium Hydroxide
Potassium hydroxide is an inorganic compound with the formula KOH , and is commonly called caustic potash. Along with sodium
hydroxide (NaOH), this colorless solid is a prototypical strong base. It has many industrial and niche applications, most of which
exploit its corrosive nature and its reactivity toward acids. Its dissolution in water is strongly exothermic.
+ −
KOH(s) → K (aq) + OH (aq)

Concentrated aqueous solutions are sometimes called potassium lyes.

Magnesium Hydroxide
Magnesium hydroxide is the inorganic compound with the chemical formula Mg(OH)
2
. Magnesium hydroxide is a common
component of antacids, such as milk of magnesia, as well as laxatives.

Figure 5.9.1 : Bottle of Antacid tablets. (CC BY 2.,5; Midnightcomm).


It is a white solid with low solubility in water. Combining a solution of many magnesium salts with basic water induces
precipitation of solid Mg(OH) . However, a weak concentration of dissociated ions can be found in solution:
2

2 + −
Mg (OH) (s) −
↽⇀
− Mg (aq) + 2 OH (aq)
2

Calcium Hydroxide
Calcium hydroxide (traditionally called slaked lime) is an inorganic compound with the chemical formula Ca(OH) . It is a 2

colorless crystal or white powder. It has many names including hydrated lime, caustic lime, builders' lime, slaked lime, cal, or
pickling lime. Calcium hydroxide is used in many applications, including food preparation. Limewater is the common name for a
saturated solution of calcium hydroxide.

5.9.2 https://chem.libretexts.org/@go/page/423656
Calcium hydroxide is relatively insoluble in water, but is large enough that its solutions are basic according to the following
reaction:
2 + −
Ca (OH) (s) −
↽⇀
− Ca (aq) + 2 OH (aq)
2

Ammonia
Ammonia is a compound of nitrogen and hydrogen with the formula NH and is a colorless gas with a characteristic pungent
3

smell. It is the active product of “smelling salts,” and can quickly revive the faint of heart and light of head. Although common in
nature and in wide use, ammonia is both caustic and hazardous in its concentrated form.

Figure 5.9.1 : Ball-and-stick model of the ammonia molecule. (Public Domain; Ben Mills).
In aqueous solution, ammonia acts as a base, acquiring hydrogen ions from H 2
O to yield ammonium and hydroxide ions:
+ −
NH (g) + H O(l) −
↽⇀
− NH (aq) + OH (aq)
3 2 4

Ammonia is also a building block for the synthesis of many pharmaceutical products and is used in many commercial cleaning
products.

Summary
A brief summary of properties of bases was given.
The properties of bases mostly contrast those of acids.
Bases have many, varied uses.

5.9: Bases- Properties and Examples is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew &
Henry Agnew.
14.3: Bases- Properties and Examples by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

5.9.3 https://chem.libretexts.org/@go/page/423656
5.10: Molecular Definitions of Acids and Bases
 Learning Objectives
Identify an Arrhenius acid and an Arrhenius base.
Identify a Brønsted-Lowry acid and a Brønsted-Lowry base.
Identify conjugate acid-base pairs in an acid-base reaction.

There are three major classifications of substances known as acids or bases. The theory developed by Svante Arrhenius in 1883, the
Arrhenius definition, states that an acid produces H+ in solution and a base produces OH-. Later, two more sophisticated and
general theories were proposed. These theories are the Brønsted-Lowry and Lewis definitions of acids and bases. This section will
cover the Arrhenius and Brønsted-Lowry theories; the Lewis theory is discussed elsewhere.

The Arrhenius Theory of Acids and Bases


In 1884, the Swedish chemist Svante Arrhenius proposed two specific classifications of compounds, termed acids and bases. When
dissolved in an aqueous solution, certain ions were released into the solution. An Arrhenius acid is a compound that increases the
concentration of H ions that are present when added to water. These H+ ions form the hydronium ion (H3O+) when they
+

combine with water molecules. This process is represented in a chemical equation by adding H2O to the reactants side.
+ −
HCl(aq) → H (aq) + Cl (aq)

In this reaction, hydrochloric acid (H C l) dissociates completely into hydrogen (H+) and chlorine (Cl-) ions when dissolved in
water, thereby releasing H+ ions into solution. Formation of the hydronium ion equation:
+ −
HCl(aq) + H O(l) → H O (aq) + Cl (aq)
2 3

An Arrhenius base is a compound that increases the concentration of OH



ions that are present when added to water. The
dissociation is represented by the following equation:
+ −
NaOH (aq) → Na (aq) + OH (aq)

In this reaction, sodium hydroxide (NaOH) disassociates into sodium (Na ) and hydroxide (OH ) ions when dissolved in water,
+ −

thereby releasing OH- ions into solution.

Arrhenius acids are substances which produce hydrogen ions in solution and Arrhenius
bases are substances which produce hydroxide ions in solution.

 Limitations to the Arrhenius Theory


The Arrhenius theory has many more limitations than the other two theories. The theory does not explain the weak base
ammonia (NH3), which in the presence of water, releases hydroxide ions into solution, but does not contain OH- itself. The
Arrhenius definition of acid and base is also limited to aqueous (i.e., water) solutions.

The Brønsted-Lowry Theory of Acids and Bases


In 1923, Danish chemist Johannes Brønsted and English chemist Thomas Lowry independently proposed new definitions for acids
and bases, ones that focus on proton transfer. A Brønsted-Lowry acid is any species that can donate a proton (H+) to another
molecule. A Brønsted-Lowry base is any species that can accept a proton from another molecule. In short, a Brønsted-Lowry
acid is a proton donor (PD), while a Brønsted-Lowry base is a proton acceptor (PA).

A Brønsted-Lowry acid is a proton donor, while a Brønsted-Lowry base is a proton


acceptor.
Let us use the reaction of ammonia in water to demonstrate the Brønsted-Lowry definitions of an acid and a base. Ammonia and
water molecules are reactants, while the ammonium ion and the hydroxide ion are products:
+ −
NH (aq) + H O(ℓ) −
↽⇀
− NH (aq) + OH (aq) (5.10.1)
3 2 4

5.10.1 https://chem.libretexts.org/@go/page/423657
What has happened in this reaction is that the original water molecule has donated a hydrogen ion to the original ammonia
molecule, which in turn has accepted the hydrogen ion. We can illustrate this as follows:

Because the water molecule donates a hydrogen ion to the ammonia, it is the Brønsted-Lowry acid, while the ammonia molecule—
which accepts the hydrogen ion—is the Brønsted-Lowry base. Thus, ammonia acts as a base in both the Arrhenius sense and the
Brønsted-Lowry sense.
Is an Arrhenius acid like hydrochloric acid still an acid in the Brønsted-Lowry sense? Yes, but it requires us to understand what
really happens when HCl is dissolved in water. Recall that the hydrogen atom is a single proton surrounded by a single electron. To
make the hydrogen ion, we remove the electron, leaving a bare proton. Do we really have bare protons floating around in aqueous
solution? No, we do not. What really happens is that the H+ ion attaches itself to H2O to make H3O+, which is called the hydronium
ion. For most purposes, H+ and H3O+ represent the same species, but writing H3O+ instead of H+ shows that we understand that
there are no bare protons floating around in solution. Rather, these protons are actually attached to solvent molecules.

 The Hydronium Ion

A proton in aqueous solution may be surrounded by more than one water molecule, leading to formulas like H 5
+
O
2
or H O
9
+

rather than H O . It is simpler, however, to use H O to represent the hydronium ion.


3
+

3
+

With this in mind, how do we define HCl as an acid in the Brønsted-Lowry sense? Consider what happens when HCl is dissolved
in H2O:
+ −
HCl(g) + H O(ℓ) → H O (aq) + Cl (aq) (5.10.2)
2 3

We can depict this process using Lewis electron dot diagrams:

Now we see that a hydrogen ion is transferred from the HCl molecule to the H2O molecule to make chloride ions and hydronium
ions. As the hydrogen ion donor, HCl acts as a Brønsted-Lowry acid; as a hydrogen ion acceptor, H2O is a Brønsted-Lowry base.
So HCl is an acid not just in the Arrhenius sense, but also in the Brønsted-Lowry sense. Moreover, by the Brønsted-Lowry
definitions, H2O is a base in the formation of aqueous HCl. So the Brønsted-Lowry definitions of an acid and a base classify the
dissolving of HCl in water as a reaction between an acid and a base—although the Arrhenius definition would not have labeled
H2O a base in this circumstance.

A Brønsted-Lowry acid is a proton (hydrogen ion) donor.


A Brønsted-Lowry base is a proton (hydrogen ion) acceptor.

5.10.2 https://chem.libretexts.org/@go/page/423657
All Arrhenius acids and bases are Brønsted-Lowry acids and bases as well. However, not all Brønsted-Lowry acids and
bases are Arrhenius acids and bases.

 Example 5.10.1

Aniline (C6H5NH2) is slightly soluble in water. It has a nitrogen atom that can accept a hydrogen ion from a water molecule,
just like the nitrogen atom in ammonia does. Write the chemical equation for this reaction and identify the Brønsted-Lowry
acid and base.

Solution
C6H5NH2 and H2O are the reactants. When C6H5NH2 accepts a proton from H2O, it gains an extra H and a positive charge and
leaves an OH− ion behind. The reaction is as follows:
+ −
C H NH (aq) + H O(ℓ) −
↽⇀
− C H NH (aq) + OH (aq)
6 5 2 2 6 5 3

Because C6H5NH2 accepts a proton, it is the Brønsted-Lowry base. The H2O molecule, because it donates a proton, is the
Brønsted-Lowry acid.

 Exercise 5.10.1

Identify the Brønsted-Lowry acid and the Brønsted-Lowry base in this chemical equation.
− 2 − +
H PO +H O −
↽⇀
− HPO +H O
2 4 2 4 3

Answer
Brønsted-Lowry acid: H2PO4-; Brønsted-Lowry base: H2O

 Exercise 5.10.2

Which of the following compounds is a Bronsted-Lowry base?


a. HCl
b. HPO42-
c. H3PO4
d. NH4+
e. CH3NH3+

Answer
A Brønsted-Lowry Base is a proton acceptor, which means it will take in an H+. This eliminates HCl, H 3
PO , NH and
4
+
4

CH NH
3
+
3
because they are Bronsted-Lowry acids. They all give away protons. In the case of HPO 2 −

4
, consider the
following equation:
2 − 3 − +
HPO (aq) + H O(l) → PO (aq) + H O (aq)
4 2 4 3

Here, it is clear that HPO42- is the acid since it donates a proton to water to make H3O+ and PO43-. Now consider the
following equation:
2 − − −
HPO (aq) + H O(l) → H PO + OH (aq)
4 2 2 4

In this case, HPO42- is the base since it accepts a proton from water to form H2PO4- and OH-. Thus, HPO42- is an acid and
base together, making it amphoteric.
Since HPO42- is the only compound from the options that can act as a base, the answer is (b) HPO42-.

5.10.3 https://chem.libretexts.org/@go/page/423657
Conjugate Acid-Base Pair
In reality, all acid-base reactions involve the transfer of protons between acids and bases. For example, consider the acid-base
reaction that takes place when ammonia is dissolved in water. A water molecule (functioning as an acid) transfers a proton to an
ammonia molecule (functioning as a base), yielding the conjugate base of water, OH , and the conjugate acid of ammonia, NH :
− +
4

This figure has three parts in two rows. In the first row, two diagrams of acid-base pairs are shown. On the left, a space filling
model of H subscript 2 O is shown with a red O atom at the center and two smaller white H atoms attached in a bent shape. Above
this model is the label “H subscript 2 O (acid)” in purple. An arrow points right, which is labeled “Remove H superscript plus.” To
the right is another space filling model with a single red O atom to which a single smaller white H atom is attached. The label in
purple above this model reads, “O H superscript negative (conjugate base).” Above both of these red and white models is an
upward pointing bracket that is labeled “Conjugate acid-base pair.” To the right is a space filling model with a central blue N atom
to which three smaller white H atoms are attached in a triangular pyramid arrangement. A label in green above reads “N H
subscript 3 (base).” An arrow labeled “Add H superscript plus” points right. To the right of the arrow is another space filling model
with a blue central N atom and four smaller white H atoms in a tetrahedral arrangement. The green label above reads “N H
subscript 3 superscript plus (conjugate acid).” Above both of these blue and white models is an upward pointing bracket that is
labeled “Conjugate acid-base pair.” The second row of the figure shows the chemical reaction, H subscript 2 O ( l ) is shown in
purple, and is labeled below in purple as “acid,” plus N H subscript 3 (a q) in green, labeled below in green as “base,” followed by
a double sided arrow arrow and O H superscript negative (a q) in purple, labeled in purple as “conjugate base,” plus N H subscript
4 superscript plus (a q)” in green, which is labeled in green as “conjugate acid.” The acid on the left side of the equation is
connected to the conjugate base on the right with a purple line. Similarly, the base on the left is connected to the conjugate acid on
the right side.
In the reaction of ammonia with water to give ammonium ions and hydroxide ions, ammonia acts as a base by accepting a proton
from a water molecule, which in this case means that water is acting as an acid. In the reverse reaction, an ammonium ion acts as an
acid by donating a proton to a hydroxide ion, and the hydroxide ion acts as a base. The conjugate acid–base pairs for this reaction
are N H /N H and H O/OH .
+

4
3 2

Figure 5.10.1 . The pairing of parent acids and bases with conjugate acids and bases.

5.10.4 https://chem.libretexts.org/@go/page/423657
Figure 5.10.1 : The Relative Strengths of Some Common Conjugate Acid-Base Pairs. The strongest acids are at the bottom left, and
the strongest bases are at the top right. The conjugate base of a strong acid is a very weak base, and, conversely, the conjugate acid
of a strong base is a very weak acid.

 Example 5.10.2

Identify the conjugate acid-base pairs in this equilibrium.


+ −
CH CO H + H O −
↽⇀
− H O + CH CO
3 2 2 3 3 2

Solution
Similarly, in the reaction of acetic acid with water, acetic acid donates a proton to water, which acts as the base. In the reverse
reaction, H O is the acid that donates a proton to the acetate ion, which acts as the base.
3
+

Once again, we have two conjugate acid-base pairs:


the parent acid and its conjugate base (C H C O H /C H
3 2

3 C O2 ) and
the parent base and its conjugate acid (H O /H O).
3
+
2

5.10.5 https://chem.libretexts.org/@go/page/423657
 Example 5.10.3

Identify the conjugate acid-base pairs in this equilibrium.


+ −
(CH ) N + H O −
↽⇀
− (CH ) NH + OH
3 3 2 3 3

Solution
One pair is H2O and OH−, where H2O has one more H+ and is the conjugate acid, while OH− has one less H+ and is the
conjugate base.
The other pair consists of (CH3)3N and (CH3)3NH+, where (CH3)3NH+ is the conjugate acid (it has an additional proton) and
(CH3)3N is the conjugate base.

 Exercise 5.10.3

Identify the conjugate acid-base pairs in this equilibrium.


− −
NH + H O ⇌ NH + OH
2 2 3

Answer
H2O (acid) and OH− (base); NH2− (base) and NH3 (acid)

5.10: Molecular Definitions of Acids and Bases is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-
Agnew & Henry Agnew.
14.4: Molecular Definitions of Acids and Bases by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

5.10.6 https://chem.libretexts.org/@go/page/423657
5.11: Reactions of Acids and Bases
 Learning Objectives
Write acid-base neutralization reactions.
Write reactions of acids with metals.
Write reactions of bases with metals.

Neutralization Reactions
The reaction that happens when an acid, such as HCl, is mixed with a base, such as NaOH:

HCl(aq) + NaOH(aq) → NaCl(aq) + H O(l)


2

When an acid and a base are combined, water and a salt are the products. Salts are ionic compounds containing a positive ion other
than H and a negative ion other than the hydroxide ion, OH . Double displacement reactions of this type are called
+ −

neutralization reactions. We can write an expanded version of this equation, with aqueous substances written in their longer form:
+ − + − + −
H (aq) + Cl (aq) + Na (aq) + OH (aq) → Na (aq) + Cl (aq) + H O(l)
2

After removing the spectator ions, we get the net ionic equation:
+ −
H (aq) + OH (aq) → H O(l)
2

When a strong acid and a strong base are combined in the proper amounts—when [H ] equals [OH ]\)—a neutral solution results
+ −

in which pH = 7. The acid and base have neutralized each other, and the acidic and basic properties are no longer present.
Salt solutions do not always have a pH of 7, however. Through a process known as hydrolysis, the ions produced when an acid and
base combine may react with the water molecules to produce a solution that is slightly acidic or basic. As a general concept, if a
strong acid is mixed with a weak base, the resulting solution will be slightly acidic. If a strong base is mixed with a weak acid, the
solution will be slightly basic.

acid-base reaction (HCl + NaOH)

Video: Equimolar (~0.01 M) and equivolume solutions of HCl and NaOH are combined to make salt water. https://youtu.be/TS-
I9KrUjB0

 Example 5.11.1: Propionic Acid + Calcium Hydroxide


Calcium propionate is used to inhibit the growth of molds in foods, tobacco, and some medicines. Write a balanced chemical
equation for the reaction of aqueous propionic acid (CH3CH2CO2H) with aqueous calcium hydroxide [Ca(OH)2].

Solution
Solutions to Example 14.5.1
Steps Reaction

5.11.1 https://chem.libretexts.org/@go/page/423658
Steps Reaction

Write the unbalanced equation.


This is a double displacement reaction, so the cations and anions CH3CH2CO2H(aq) + Ca(OH)2(aq)→(CH3CH2CO2)2Ca(aq) + H2O(l)
swap to create the water and the salt.

Balance the equation.


2CH3CH2CO2H(aq) + Ca(OH)2(aq)→(CH3CH2CO2)2Ca(aq)
Because there are two OH− ions in the formula for Ca(OH)2, we need
+2H2O(l)
two moles of propionic acid, CH3CH2CO2H, to provide H+ ions.

 Exercise 5.11.1

Write a balanced chemical equation for the reaction of solid barium hydroxide with dilute acetic acid.

Answer
Ba (OH) (s) + 2 CH CO H(aq) → Ba (CH CO ) (aq) + 2 H O(l)
2 3 2 3 2 2 2

Acids and Bases React with Metals


Acids react with most metals to form a salt and hydrogen gas. As discussed previously, metals that are more active than acids can
undergo a single displacement reaction. For example, zinc metal reacts with hydrochloric acid, producing zinc chloride and
hydrogen gas.

Zn(s) + 2 HCl(aq) → ZnCl (aq) + H (g)


2 2

Bases also react with certain metals, like zinc or aluminum, to produce hydrogen gas. For example, sodium hydroxide reacts with
zinc and water to form sodium zincate and hydrogen gas.

Zn(s) + 2 NaOH(aq) + 2 H O(l) → Na Zn(OH) (aq) + H (g).


2 2 4 2

5.11: Reactions of Acids and Bases is shared under a Public Domain license and was authored, remixed, and/or curated by Marisa Alviar-Agnew,
Henry Agnew, Peggy Lawson, & Peggy Lawson.
14.5: Reactions of Acids and Bases by Henry Agnew, Marisa Alviar-Agnew, Peggy Lawson is licensed Public Domain. Original source:
https://sites.prairiesouth.ca/legacy/chemistry/chem30.

5.11.2 https://chem.libretexts.org/@go/page/423658
5.12: Strong and Weak Acids and Bases
 Learning Objectives
Define a strong and a weak acid and base.
Recognize an acid or a base as strong or weak.
Determine if a salt produces an acidic or a basic solution.

Strong and Weak Acids


Except for their names and formulas, so far we have treated all acids as equals, especially in a chemical reaction. However, acids can be very different in a very important way. Consider HCl(aq).
When HCl is dissolved in H2O, it completely dissociates into H+(aq) and Cl−(aq) ions; all the HCl molecules become ions:
100%
+ −
HC l → H (aq) + C l (aq)

Any acid that dissociates 100% into ions is called a strong acid. If it does not dissociate 100%, it is a weak acid. HC2H3O2 is an example of a weak acid:
∼5%
+ −
H C2 H3 O2 ⟶ H (aq) + C2 H3 O (aq)
2

Because this reaction does not go 100% to completion, it is more appropriate to write it as a reversible reaction:
+ −
H C2 H3 O2 ⇌ H (aq) + C2 H3 O (aq)
2

As it turns out, there are very few strong acids, which are given in Table 5.12.1. If an acid is not listed here, it is a weak acid. It may be 1% ionized or 99% ionized, but it is still classified as a weak
acid.

Any acid that dissociates 100% into ions is called a strong acid. If it does not dissociate 100%, it is a weak acid.
Table 5.12.1 : Strong Acids and Bases
Acids Bases

HCl LiOH

HBr NaOH

HI KOH

HNO3 RbOH

H2SO4 CsOH

HClO3 Mg(OH)2

HClO4 Ca(OH)2

Sr(OH)2

Ba(OH)2

Strong and Weak Bases


The issue is similar with bases: a strong base is a base that is 100% ionized in solution. If it is less than 100% ionized in solution, it is a weak base. There are very few strong bases (Table 5.12.1);
any base not listed is a weak base. All strong bases are OH– compounds. So a base based on some other mechanism, such as NH3 (which does not contain OH− ions as part of its formula), will be a
weak base.

 Example 5.12.1: Identifying Strong and Weak Acids and Bases

Identify each acid or base as strong or weak.


a. HCl
b. Mg(OH)2
c. C5H5N

Solution
a. Because HCl is listed in Table 5.12.1, it is a strong acid.
b. Because Mg(OH)2 is listed in Table 5.12.1, it is a strong base.
c. The nitrogen in C5H5N would act as a proton acceptor and therefore can be considered a base, but because it does not contain an OH compound, it cannot be considered a strong base; it is a
weak base.

 Exercise 5.12.1

Identify each acid or base as strong or weak.


a. RbOH
b. HNO 2

Answer a
strong base
Answer b
weak acid

 Example 5.12.2: Characterizing Base Ionization

Write the balanced chemical equation for the dissociation of Ca(OH)2 and indicate whether it proceeds 100% to products or not.

Solution
This is an ionic compound of Ca2+ ions and OH− ions. When an ionic compound dissolves, it separates into its constituent ions:

5.12.1 https://chem.libretexts.org/@go/page/423659
2 + −
Ca (OH) → Ca (aq) + 2 OH (aq)
2

Because Ca(OH)2 is listed in Table 5.12.1, this reaction proceeds 100% to products.

 Exercise 5.12.2

Write the balanced chemical equation for the dissociation of hydrazoic acid (HN3) and indicate whether it proceeds 100% to products or not.

Answer a
The reaction is as follows:
+ −
HN → H (aq) + N (aq)
3 3

It does not proceed 100% to products because hydrazoic acid is not a strong acid.

Key Takeaways
Strong acids and bases are 100% ionized in aqueous solution.
Weak acids and bases are less than 100% ionized in aqueous solution.
Salts of weak acids or bases can affect the acidity or basicity of their aqueous solutions.

Contributions & Attributions

5.12: Strong and Weak Acids and Bases is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew & Henry Agnew.
14.7: Strong and Weak Acids and Bases by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source: https://www.ck12.org/c/chemistry/.

5.12.2 https://chem.libretexts.org/@go/page/423659
5.13: Acid–Base Titration
 Learning Objectives
Understand the basics of acid-base titrations.
Understand the use of indicators.
Perform a titration calculation correctly.

The reaction of an acid with a base to make a salt and water is a common reaction in the laboratory, partly because so many
compounds can act as acids or bases. Another reason that acid-base reactions are so prevalent is because they are often used to
determine quantitative amounts of one or the other. Performing chemical reactions quantitatively to determine the exact amount of
a reagent is called a titration. A titration can be performed with almost any chemical reaction for which the balanced chemical
equation is known. Here, we will consider titrations that involve acid-base reactions.
During an acid-base titration, an acid with a known concentration (a standard solution) is slowly added to a base with an unknown
concentration (or vice versa). A few drops of indicator solution are added to the base. The indicator will signal, by color change,
when the base has been neutralized (when [H+] = [OH-]). At that point—called the equivalence point, or end point—the titration
is stopped. By knowing the volumes of acid and base used, and the concentration of the standard solution, calculations allow us to
determine the concentration of the other solution.

It is important to accurately measure volumes when doing titrations. The instrument you would use is called a burette (or
buret).

Figure 5.13.1 : Equipment for Titrations. A burette is a type of liquid dispensing system that can accurately indicate the volume of
liquid dispensed.
For example, suppose 25.66 mL (or 0.02566 L) of 0.1078 M HCl was used to titrate an unknown sample of NaOH. What mass of
NaOH was in the sample? We can calculate the number of moles of HCl reacted:
# mol HCl = (0.02566 L)(0.1078 M) = 0.002766 mol HCl
We also have the balanced chemical reaction between HCl and NaOH:

HCl + NaOH → NaCl + H O


2

So we can construct a conversion factor to convert to number of moles of NaOH reacted:


1 mol N aOH
0.002766 mol H C l × = 0.002766 mol N aOH
1 mol H C l

Then we convert this amount to mass, using the molar mass of NaOH (40.00 g/mol):
40.00 g N aOH
0.002766 mol H C l × = 0.1106 g N aOH
1 mol H C l

This type of calculation is performed as part of a titration.

 Example 5.13.1: Equivalence Point


What mass of Ca(OH)2 is present in a sample if it is titrated to its equivalence point with 44.02 mL of 0.0885 M HNO3? The
balanced chemical equation is as follows:

2 HNO + Ca (OH) → Ca (NO ) +2 H O


3 2 3 2 2

5.13.1 https://chem.libretexts.org/@go/page/423660
Solution
In liters, the volume is 0.04402 L. We calculate the number of moles of titrant:
# moles HNO3 = (0.04402 L)(0.0885 M) = 0.00390 mol HNO3
Using the balanced chemical equation, we can determine the number of moles of Ca(OH)2 present in the analyte:
1 mol C a(OH )2
0.00390 mol H N O3 × = 0.00195 mol C a(OH )2
2 mol H N O3

Then we convert this to a mass using the molar mass of Ca(OH)2:


74.1 g C a(OH )2
0.00195 mol C a(OH )2 × = 0.144 g C a(OH )2
mol C a(OH )2

 Exercise 5.13.1

What mass of H2C2O4 is present in a sample if it is titrated to its equivalence point with 18.09 mL of 0.2235 M NaOH? The
balanced chemical reaction is as follows:

H C O + 2 NaOH → Na C O +2 H O
2 2 4 2 2 4 2

Answer
0.182 g

 Exercise 5.13.2

If 25.00 mL of HCl solution with a concentration of 0.1234 M is neutralized by 23.45 mL of NaOH, what is the concentration
of the base?

Answer
0.1316 M NaOH

 Exercise 5.13.3

A 20.0 mL solution of strontium hydroxide, Sr(OH)2, is placed in a flask and a drop of indicator is added. The solution turns
color after 25.0 mL of a standard 0.0500 M HCl solution is added. What was the original concentration of the Sr(OH)2
solution?

Answer
3.12 × 10
−2
M Sr(OH)2

Indicator Selection for Titrations


The indicator used depends on the type of titration performed. The indicator of choice should change color when enough of one
substance (acid or base) has been added to exactly use up the other substance. Only when a strong acid and a strong base are
produced will the resulting solution be neutral. The three main types of acid-base titrations, and suggested indicators, are:
The three main types of acid-base titrations, suggested indicators, and explanations
Titration between . . . Indicator Explanation

strong acid and strong base any

strong acid and weak base methyl orange changes color in the acidic range (3.2 - 4.4)

weak acid and strong base phenolphthalein changes color in the basic range (8.2 - 10.6)

5.13.2 https://chem.libretexts.org/@go/page/423660
Summary
A titration is the quantitative reaction of an acid and a base. Indicators are used to show that all the analyte has reacted with the
titrant.

Contributions & Attributions


Peggy Lawson (Oxbow Prairie Heights School). Funded by Saskatchewan Educational Technology Consortium.

5.13: Acid–Base Titration is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew & Henry Agnew.
14.6: Acid–Base Titration by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source: https://www.ck12.org/c/chemistry/.

5.13.3 https://chem.libretexts.org/@go/page/423660
5.14: Water - Acid and Base in One
 Learning Objectives
Describe the autoionization of water.
Calculate the concentrations of H O and OH in aqueous solutions, knowing the other concentration.
3
+ −

We have already seen that H 2


O can act as an acid or a base:
+ −
NH +H O −
↽⇀
− NH + OH
3 2 4
 
base acid

where H 2
O acts as an acid (in red).
+ −
HCl + H O ⟶ H O + Cl
2 3


acid
base

where H 2
O acts as an base (in blue).
It may not surprise you to learn, then, that within any given sample of water, some H O
2
molecules are acting as acids, and other
H O molecules are acting as bases. The chemical equation is as follows:
2

+ −
H O+H O −
↽⇀
− H O + OH (5.14.1)
2 2 3
 
acid base

This occurs only to a very small degree: only about 6 in 108 H 2


O molecules are participating in this process, which is called the
autoionization of water.

Figure 5.14.1 : Autoionization of water, resulting in hydroxide and hydronium ions.


At this level, the concentration of both H O (aq) and OH (aq) in a sample of pure H O is about 1.0 × 10 M (at room
3
+ −

2
−7

temperature). If we use square brackets—[ ]—around a dissolved species to imply the molar concentration of that species, we have
+ − −7
[H O ]=[ OH ] = 1.0 × 10 (5.14.2)
3

for any sample of pure water because H2O can act as both an acid and a base. The product of these two concentrations is
1.0 × 10 :
−14

+ − −7 −7 −14
[H O ]×[ OH ] = (1.0 × 10 )(1.0 × 10 ) = 1.0 × 10
3

For acids, the concentration of H O (aq) (i.e., [H O ]) is greater than 1.0 × 10 M .


3
+

3
+ −7

For bases the concentration of OH (aq) (i.e., [OH ]) is greater than 1.0 × 10 M .
− − −7

However, the product of the two concentrations—[H 3


O
+
][ OH

] —is always equal to 1.0 × 10 −14
, no matter whether the aqueous
solution is an acid, a base, or neutral:
+ − −14
[H O ][ OH ] = 1.0 × 10
3

This value of the product of concentrations is so important for aqueous solutions that it is called the autoionization constant of
water and is denoted K : w

+ − −14
Kw = [ H O ][ OH ] = 1.0 × 10 (5.14.3)
3

This means that if you know [H O ] for a solution, you can calculate what [OH ]) has to be for the product to equal 1.0 × 10 ;
3
+ − −14

or if you know [OH ]), you can calculate [H O ]. This also implies that as one concentration goes up, the other must go down to

3
+

compensate so that their product always equals the value of K . w

5.14.1 https://chem.libretexts.org/@go/page/423661
 Warning: Temperature Matters

The degree of autoionization of water (Equation 5.14.1)—and hence the value of Kw —changes with temperature, so
Equations 5.14.2 - 5.14.3 are accurate only at room temperature.

 Example 5.14.1: Hydroxide Concentration

What is [OH ]) of an aqueous solution if [H


3
+
O ] is 1.0 × 10 −4
M ?

Solution
Solutions to Example 14.7.1
Steps for Problem Solving

Identify the "given" information and what the problem is asking you Given: [H O ] = 1.0 × 10
3
+ −4
M

to "find." Find: [OH−] = ? M

List other known quantities. none

Using the expression for K , (Equation


w 5.14.3 ), rearrange the
equation algebraically to solve for [OH−].
Plan the problem. −14
1.0 × 10

[ OH ] =
+
[H3 O ]

Now substitute the known quantities into the equation and solve.
−14
1.0 × 10
− −10
[ OH ] = = 1.0 × 10 M
Calculate. 1.0 × 10
−4

It is assumed that the concentration unit is molarity, so [OH



] is 1.0
× 10−10 M.
The concentration of the acid is high (> 1 x 10-7 M), so [OH −
]
Think about your result.
should be low.

 Exercise 5.14.1

What is [OH −
] in a 0.00032 M solution of H2SO4?

Hint
Assume both protons ionize from the molecule...although this is not the case.

Answer
−11
3.1 × 10 M

When you have a solution of a particular acid or base, you need to look at the formula of the acid or base to determine the number
of H3O+ or OH− ions in the formula unit because [H O ] or [OH ]) may not be the same as the concentration of the acid or base
3
+ −

itself.

 Example 5.14.2: Hydronium Concentration


What is [H 3
O
+
] in a 0.0044 M solution of Ca(OH) ? 2

Solution
Solutions to Example 14.7.2

5.14.2 https://chem.libretexts.org/@go/page/423661
Steps for Problem Solving

Identify the "given" information and what the problem is asking you Given: [Ca(OH) ] = 0.0044 M2

to "find." Find: [H O ] = ? M
3
+

We begin by determining [OH ] . The concentration of the solute is


0.0044 M, but because Ca(OH) is a strong base, there are two OH−
2

ions in solution for every formula unit dissolved, so the actual [OH ] −

List other known quantities.


is two times this:

[OH ]=2 × 0 ⋅ 0044 M=0 ⋅ 0088 M⋅

Use the expression for K (Equation 5.14.3 ) and rearrange the


w

equation algebraically to solve for [H O ]. 3


+

Plan the problem. −14


1.0 × 10
+
[ H3 O ] =

[O H ]

Now substitute the known quantities into the equation and solve.
−14
1.0 × 10
Calculate. [ H3 O
+
] = = 1.1 × 10
−12
M
(0.0088)

[H O
3
+
] has decreased significantly in this basic solution.
The concentration of the base is high (> 1 x 10-7 M) so [H +
O ]
Think about your result. 3

should be low.

 Exercise 5.14.2

What is [H 3
+
O ] of an aqueous solution if [OH −
] is 1.0 × 10 −9
M ?

Answer
1.0 × 10−5 M

In any aqueous solution, the product of [H 3


+
O ] and [OH −
] equals 1.0 × 10 −14
(at room temperature).

Contributions & Attributions


5.14: Water - Acid and Base in One is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-Agnew & Henry
Agnew.
14.8: Water - Acid and Base in One by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

5.14.3 https://chem.libretexts.org/@go/page/423661
5.15: The pH and pOH Scales - Ways to Express Acidity and Basicity
 Learning Objectives
Define pH and pOH.
Determine the pH of acidic and basic solutions.
Determine the hydronium ion concentration and pOH from pH.

As we have seen, [H O ] and [OH ] values can be markedly different from one aqueous solution to another. So chemists defined
3
+ −

a new scale that succinctly indicates the concentrations of either of these two ions.
pH is a logarithmic function of [H 3O
+
:
]

+
pH = − log[ H3 O ] (5.15.1)

pH is usually (but not always) between 0 and 14. Knowing the dependence of pH on [H 3O
+
, we can summarize as follows:
]

If pH < 7, then the solution is acidic.


If pH = 7, then the solution is neutral.
If pH > 7, then the solution is basic.
This is known as the pH scale. The pH scale is the range of values from 0 to 14 that describes the acidity or basicity of a solution.
You can use pH to make a quick determination whether a given aqueous solution is acidic, basic, or neutral. Figure 5.15.1
illustrates this relationship, along with some examples of various solutions. Because hydrogen ion concentrations are generally less
than one (for example 1.3 × 10 M ), the log of the number will be a negative number. To make pH even easier to work with, pH
−3

is defined as the negative log of [H O ], which will give a positive value for pH.
3
+

Figure 5.15.1 : The pH values for several common materials.

 Example 5.15.1

Label each solution as acidic, basic, or neutral based only on the stated pH.
a. milk of magnesia, pH = 10.5
b. pure water, pH = 7
c. wine, pH = 3.0

Answer

5.15.1 https://chem.libretexts.org/@go/page/423662
a. With a pH greater than 7, milk of magnesia is basic. (Milk of magnesia is largely Mg(OH)2.)
b. Pure water, with a pH of 7, is neutral.
c. With a pH of less than 7, wine is acidic.

 Exercise 5.15.1

Identify each substance as acidic, basic, or neutral based only on the stated pH.
a. human blood with pH = 7.4
b. household ammonia with pH = 11.0
c. cherries with pH = 3.6

Answer a
basic

Answer b
basic

Answer c
acidic

Calculating pH from Hydronium Concentration


The pH of solutions can be determined by using logarithms as illustrated in the next example for stomach acid. Stomach acid is a
solution of H C l with a hydronium ion concentration of 1.2 × 10 M , what is the pH of the solution?
−3

+
pH = − log[ H3 O ]

−3
= − log(1.2 × 10 )

= −(−2.92) = 2.92

 Logarithms

To get the log value on your calculator, enter the number (in this case, the hydronium ion concentration) first, then press the
LOG key.
If the number is 1.0 x 10-5 (for [H3O+] = 1.0 x 10-5 M) you should get an answer of "-5".
If you get a different answer, or an error, try pressing the LOG key before you enter the number.

 Example 5.15.2: Converting Ph to Hydronium Concentration


Find the pH, given the [H 3O
+
] of the following:
a. 1 ×10-3 M
b. 2.5 ×10-11 M
c. 4.7 ×10-9 M

Solution
Solutions to Example 14.9.2
Steps for Problem Solving

Given:
a. [H3O+] =1 × 10−3 M
Identify the "given" information and what the problem is asking you
b. [H3O+] =2.5 ×10-11 M
to "find."
c. [H3O+] = 4.7 ×10-9 M
Find: ? pH

5.15.2 https://chem.libretexts.org/@go/page/423662
Steps for Problem Solving

Need to use the expression for pH (Equation 5.15.1 ).


Plan the problem.
pH = - log [H3O+]
Now substitute the known quantity into the equation and solve.
a. pH = - log [1 × 10−3 ] = 3.0 (1 decimal places since 1 has 1
significant figure)
b. pH = - log [2.5 ×10-11] = 10.60 (2 decimal places since 2.5 has 2
significant figures)
c. pH = - log [4.7 ×10-9] = 8.30 (2 decimal places since 4.7 has 2
significant figures)
Calculate.
The other issue that concerns us here is significant figures. Because
the number(s) before the decimal point in a logarithm relate to the
power on 10, the number of digits after the decimal point is what
determines the number of significant figures in the final answer:

 Exercise 5.15.2

Find the pH, given [H3O+] of the following:


a. 5.8 ×10-4 M
b. 1.0×10-7

Answer a
3.22

Answer b
7.00

Calculating Hydronium Concentration from pH


Sometimes you need to work "backwards"—you know the pH of a solution and need to find [H O ], or even the concentration of
3
+

the acid solution. How do you do that? To convert pH into [H O ] we solve Equation 5.15.1 for [H O ]. This involves taking the
3
+
3
+

antilog (or inverse log) of the negative value of pH .


+
[H O ] = antilog(−pH )
3

or
+ −pH
[H O ] = 10 (5.15.2)
3

As mentioned above, different calculators work slightly differently—make sure you can do the following calculations using your
calculator.

 Calculator Skills

We have a solution with a pH = 8.3. What is [H3O+] ?

With some calculators you will do things in the following order:


1. Enter 8.3 as a negative number (use the key with both the +/- signs, not the subtraction key).
2. Use your calculator's 2nd or Shift or INV (inverse) key to type in the symbol found above the LOG key. The shifted
function should be 10x.

5.15.3 https://chem.libretexts.org/@go/page/423662
3. You should get the answer 5.0 × 10-9.

Other calculators require you to enter keys in the order they appear in the equation.
1. Use the Shift or second function to key in the 10x function.
2. Use the +/- key to type in a negative number, then type in 8.3.
3. You should get the answer 5.0 × 10-9.
If neither of these methods work, try rearranging the order in which you type in the keys. Don't give up—you must master your
calculator!

 Example 5.15.3: Calculating Hydronium Concentration from pH

Find the hydronium ion concentration in a solution with a pH of 12.6. Is this solution an acid or a base? How do you know?

Solution
Solutions to Example 14.9.3
Steps for Problem Solving

Identify the "given" information and what the problem is asking you Given: pH = 12.6
to "find." Find: [H3O+] = ? M

Need to use the expression for [H3O+] (Equation 5.15.2 ).


Plan the problem.
[H3O+] = antilog (-pH) or [H3O+] = 10-pH
Now substitute the known quantity into the equation and solve.
[H3O+] = antilog (12.60) = 2.5 x 10-13 M (2 significant figures since
4.7 has 12.60 2 decimal places)
or
[H3O+] = 10-12.60 = 2.5 x 10-13 M (2 significant figures since 4.7 has
12.60 2 decimal places)
Calculate. The other issue that concerns us here is significant figures. Because
the number(s) before the decimal point in a logarithm relate to the
power on 10, the number of digits after the decimal point is what
determines the number of significant figures in the final answer:

 Exercise 5.15.3

If moist soil has a pH of 7.84, what is [H3O+] of the soil solution?

Answer
1.5 x 10-8 M

The pOH scale


As with the hydrogen-ion concentration, the concentration of the hydroxide ion can be expressed logarithmically by the pOH. The
pOH of a solution is the negative logarithm of the hydroxide-ion concentration.

pOH = −log [ OH ]

The relation between the hydronium and hydroxide ion concentrations expressed as p-functions is easily derived from the Kw

expression:

5.15.4 https://chem.libretexts.org/@go/page/423662
+ −
Kw = [ H O ][ OH ] (5.15.3)
3

+ − + −
− log Kw = − log([ H3 O ][OH ]) = − log[ H3 O ] + − log[OH ] (5.15.4)

pKw = pH + pOH (5.15.5)

At 25 °C, the value of K is 1.0 × 10


w
−14
, and so:
14.00 = pH + pOH (5.15.6)

The hydronium ion molarity in pure water (or any neutral solution) is 1.0 × 10
−7
M at 25 °C. The pH and pOH of a neutral
solution at this temperature are therefore:
+ −7
pH = − log[ H3 O ] = − log(1.0 × 10 ) = 7.00 (5.15.7)

− −7
pOH = − log[OH ] = − log(1.0 × 10 ) = 7.00 (5.15.8)

And so, at this temperature, acidic solutions are those with hydronium ion molarities greater than 1.0 × 10 M and hydroxide ion −7

molarities less than 1.0 × 10 M (corresponding to pH values less than 7.00 and pOH values greater than 7.00). Basic solutions
−7

are those with hydronium ion molarities less than 1.0 × 10 M and hydroxide ion molarities greater than 1.0 × 10 M
−7 −7

(corresponding to pH values greater than 7.00 and pOH values less than 7.00).

 Example 5.15.4:

Find the pOH of a solution with a pH of 4.42.

Solution
Solutions to Example 14.9.4
Steps for Problem Solving

Identify the "given" information and what the problem is asking you Given: pH =4.42
to "find." Find: ? pOH

Need to use the expression


Plan the problem.
pOH = 14 - pH
Now substitute the known quantity into the equation and solve.
Calculate.
pOH=14−4.42=9.58
The pH is that of an acidic solution, and the resulting pOH is the
Think about your result. difference after subtracting from 14. The answer has two significant
figures because the given pH has two decimal places.

 Exercise 5.15.4

The pH of a solution is 8.22. What is the pOH?

Answer
5.78

The diagram below shows all of the interrelationships between [H3O+][H3O+], [OH−][OH−], pH, and pOH.

5.15.5 https://chem.libretexts.org/@go/page/423662
Figure 5.15.1 : Relationships between hydrogen ion concentration, hydroxide ion concentration, pH and pOH.

Contributions & Attributions


Peggy Lawson (Oxbow Prairie Heights School). Funded by Saskatchewan Educational Technology Consortium.

Template:OpenStax

5.15: The pH and pOH Scales - Ways to Express Acidity and Basicity is shared under a CK-12 license and was authored, remixed, and/or curated
by Marisa Alviar-Agnew & Henry Agnew.
14.9: The pH and pOH Scales - Ways to Express Acidity and Basicity by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original
source: https://www.ck12.org/c/chemistry/.

5.15.6 https://chem.libretexts.org/@go/page/423662
5.16: Buffers- Solutions that Resist pH Change
 Learning Objective
Define buffer and describe how it reacts with an acid or a base.

Weak acids are relatively common, even in the foods we eat. But we occasionally come across a strong acid or base, such as
stomach acid, that has a strongly acidic pH of 1–2. By definition, strong acids and bases can produce a relatively large amount of
hydrogen or hydroxide ions and, as a consequence, have marked chemical activity. In addition, very small amounts of strong acids
and bases can change the pH of a solution very quickly. If 1 mL of stomach acid [which we will approximate as 0.05 M HCl(aq)] is
added to the bloodstream, and if no correcting mechanism is present, the pH of the blood would go from about 7.4 to about 4.9—a
pH that is not conducive to life. Fortunately, the body has a mechanism for minimizing such dramatic pH changes.
This mechanism involves a buffer, a solution that resists dramatic changes in pH. Buffers do so by being composed of certain pairs
of solutes: either a weak acid plus a salt derived from that weak acid, or a weak base plus a salt of that weak base. For example, a
buffer can be composed of dissolved acetic acid (HC2H3O2, a weak acid) and sodium acetate (NaC2H3O2, a salt derived from that
acid). Another example of a buffer is a solution containing ammonia (NH3, a weak base) and ammonium chloride (NH4Cl, a salt
derived from that base).
Let us use an acetic acid–sodium acetate buffer to demonstrate how buffers work. If a strong base—a source of OH −
(aq) ions—is
added to the buffer solution, those hydroxide ions will react with the acetic acid in an acid-base reaction:
− −
HC H O (aq) + OH (aq) → H O(ℓ) + C H O2 (aq) (5.16.1)
2 3 2 2 2 3

Rather than changing the pH dramatically by making the solution basic, the added hydroxide ions react to make water, and the pH
does not change much.

Many people are aware of the concept of buffers from buffered aspirin, which is aspirin that also has magnesium carbonate,
calcium carbonate, magnesium oxide, or some other salt. The salt acts like a base, while aspirin is itself a weak acid.

If a strong acid—a source of H+ ions—is added to the buffer solution, the H+ ions will react with the anion from the salt. Because
HC2H3O2 is a weak acid, it is not ionized much. This means that if lots of hydrogen ions and acetate ions (from sodium acetate) are
present in the same solution, they will come together to make acetic acid:
+ −
H (aq) + C H O2 (aq) → HC H O (aq) (5.16.2)
2 3 2 3 2

Rather than changing the pH dramatically and making the solution acidic, the added hydrogen ions react to make molecules of a
weak acid. Figure 5.16.1 illustrates both actions of a buffer.

Figure 5.16.1 : The Action of Buffers. Buffers can react with both strong acids (top) and strong bases (bottom) to minimize large
changes in pH.
Buffers made from weak bases and salts of weak bases act similarly. For example, in a buffer containing NH3 and NH4Cl, ammonia
molecules can react with any excess hydrogen ions introduced by strong acids:
+ +
NH (aq) + H (aq) → NH (aq) (5.16.3)
3 4

while the ammonium ion (NH +


4
(aq) ) can react with any hydroxide ions introduced by strong bases:
+ −
NH (aq) + OH (aq) → NH (aq) + H O(ℓ) (5.16.4)
4 3 2

5.16.1 https://chem.libretexts.org/@go/page/423663
 Example 5.16.1: Making Buffer Solutions

Which solute combinations can make a buffer solution? Assume that all are aqueous solutions.
a. HCHO2 and NaCHO2
b. HCl and NaCl
c. CH3NH2 and CH3NH3Cl
d. NH3 and NaOH

Solution
a. Formic acid (HCHO2) is a weak acid, while NaCHO2 is the salt made from the anion of the weak acid—the formate ion
(CHO2−). The combination of these two solutes would make a buffer solution.
b. Hydrochloric acid (HCl) is a strong acid, not a weak acid, so the combination of these two solutes would not make a buffer
solution.
c. Methylamine (CH3NH2) is like ammonia with one of its hydrogen atoms substituted with a CH3 (methyl) group. Because it
is not on our list of strong bases, we can assume that it is a weak base. The compound CH3NH3Cl is a salt made from that
weak base, so the combination of these two solutes would make a buffer solution.
d. Ammonia (NH3) is a weak base, but NaOH is a strong base. The combination of these two solutes would not make a buffer
solution.

 Exercise 5.16.1

Which solute combinations can make a buffer solution? Assume that all are aqueous solutions.
a. NaHCO3 and NaCl
b. H3PO4 and NaH2PO4
c. NH3 and (NH4)3PO4
d. NaOH and NaCl

Answer a
Yes.

Answer b
No. Need a weak acid or base and a salt of its conjugate base or acid.

Answer c
Yes.

Answer d
No. Need a weak base or acid.

Buffers work well only for limited amounts of added strong acid or base. Once either solute is all reacted, the solution is no longer
a buffer, and rapid changes in pH may occur. We say that a buffer has a certain capacity. Buffers that have more solute dissolved in
them to start with have larger capacities, as might be expected.
Human blood has a buffering system to minimize extreme changes in pH. One buffer in blood is based on the presence of HCO3−
and H2CO3 [H2CO3 is another way to write CO2(aq)]. With this buffer present, even if some stomach acid were to find its way
directly into the bloodstream, the change in the pH of blood would be minimal. Inside many of the body’s cells, there is a buffering
system based on phosphate ions.

5.16.2 https://chem.libretexts.org/@go/page/423663
 Career Focus: Blood Bank Technology Specialist

At this point in this text, you should have the idea that the chemistry of blood is fairly complex. Because of this, people who
work with blood must be specially trained to work with it properly.
A blood bank technology specialist is trained to perform routine and special tests on blood samples from blood banks or
transfusion centers. This specialist measures the pH of blood, types it (according to the blood’s ABO+/− type, Rh factors, and
other typing schemes), tests it for the presence or absence of various diseases, and uses the blood to determine if a patient has
any of several medical problems, such as anemia. A blood bank technology specialist may also interview and prepare donors to
give blood and may actually collect the blood donation.
Blood bank technology specialists are well trained. Typically, they require a college degree with at least a year of special
training in blood biology and chemistry. In the United States, training must conform to standards established by the American
Association of Blood Banks.

Key Takeaway
A buffer is a solution that resists sudden changes in pH.

Contributions & Attributions

5.16: Buffers- Solutions that Resist pH Change is shared under a CK-12 license and was authored, remixed, and/or curated by Marisa Alviar-
Agnew & Henry Agnew.
14.10: Buffers- Solutions that Resist pH Change by Henry Agnew, Marisa Alviar-Agnew is licensed CK-12. Original source:
https://www.ck12.org/c/chemistry/.

5.16.3 https://chem.libretexts.org/@go/page/423663
Index
A dilution P
Acid–Base Titration 5.5: Solution Dilution pascal (unit)
5.13: Acid–Base Titration Dimensional Analysis 4.2: Pressure - The Result of Constant Molecular
ammonium ion 1.19: Problem Solving and Unit Conversions Collisions
5.16: Buffers- Solutions that Resist pH Change pH scale
anion E 5.15: The pH and pOH Scales - Ways to Express
electromagnetic spectrum Acidity and Basicity
2.7: Ions - Losing and Gaining Electrons
atomic mass unit 2.11: The Electromagnetic Spectrum Physical change
electron 1.8: Changes in Matter - Physical and Chemical
2.3: The Properties of Protons, Neutrons, and
Changes
Electrons 2.3: The Properties of Protons, Neutrons, and
atomic theory Electrons physical property
electronegativity 1.7: Differences in Matter- Physical and Chemical
2.1: Indivisible - The Atomic Theory
Properties
autoionization of water 2.23: Electronegativity and Polarity - Why Oil and
pOH
Water Do not Mix
5.14: Water - Acid and Base in One
element 5.15: The pH and pOH Scales - Ways to Express
Acidity and Basicity
B 1.6: Classifying Matter According to Its
Composition polarity
balanced chemical equation endothermic process 2.23: Electronegativity and Polarity - Why Oil and
3.10: Making Molecules- Mole-to-Mole Conversions Water Do not Mix
1.13: Energy and Chemical and Physical Change
Balancing a Chemical Equation Potential Energy
evaporation
3.8: How to Write Balanced Chemical Equations 1.12: Energy
4.12: Evaporation and Condensation
Bohr model proton
Exceptions to the Octet Rule
2.12: The Bohr Model - Atoms with Orbits 2.3: The Properties of Protons, Neutrons, and
2.21: Writing Lewis Structures for Covalent Electrons
Boyle’s law Compounds
4.3: Boyle’s Law - Pressure and Volume exothermic process
buffer 1.13: Energy and Chemical and Physical Change
R
5.16: Buffers- Solutions that Resist pH Change relative abundances
buffer capacity F 2.9: Atomic Mass - The Average Mass of an
Element’s Atoms
5.16: Buffers- Solutions that Resist pH Change formula unit rounding
3.12: Stoichiometry 1.17: Significant Figures in Calculations
C freezing
cation 4.13: Melting, Freezing, and Sublimation S
2.7: Ions - Losing and Gaining Electrons
scienfitic notation
Charles's Law H 1.15: Scientific Notation - Writing Large and Small
4.4: Charles’s Law- Volume and Temperature hydrolysis Numbers
chemical change 5.11: Reactions of Acids and Bases Separation of Mixtures
1.8: Changes in Matter - Physical and Chemical 1.8: Changes in Matter - Physical and Chemical
Changes
I Changes
chemical property significant figures
1.7: Differences in Matter- Physical and Chemical
ions
2.7: Ions - Losing and Gaining Electrons 1.16: Significant Figures - Writing Numbers to
Properties Reflect Precision
coefficient isotopes 1.17: Significant Figures in Calculations
3.10: Making Molecules- Mole-to-Mole Conversions 2.8: Isotopes - When the Number of Neutrons Varies solute
Coefficients and Subscripts 5.1: Solutions - Homogeneous Mixtures
3.8: How to Write Balanced Chemical Equations M Solution Stoichiometry
compound matter 5.6: Solution Stoichiometry
1.6: Classifying Matter According to Its 1.3: What is Matter? solvent
Composition melting 5.1: Solutions - Homogeneous Mixtures
condensation 4.13: Melting, Freezing, and Sublimation Stability of Isotopes
4.12: Evaporation and Condensation melting point 2.8: Isotopes - When the Number of Neutrons Varies
Conservation of Energy 4.9: Interactions between Molecules Stock system
1.13: Energy and Chemical and Physical Change mixture 2.27: Naming Ionic Compounds
conservation of mass 1.6: Classifying Matter According to Its stoichiometry
1.9: Conservation of Mass - There is No New Matter Composition
3.12: Stoichiometry
crisscross method molarity sublimation
2.25: Writing Formulas for Ionic Compounds 5.4: Specifying Solution Concentration- Molarity
4.13: Melting, Freezing, and Sublimation
mole ratio surface tension
D 3.12: Stoichiometry
4.11: Intermolecular Forces in Action- Surface
Dalton’s law of partial pressures Tension and Viscosity
4.8: Mixtures of Gases - Why Deep-Sea Divers N
Breathe a Mixture of Helium and Oxygen neutralization reaction U
Democritus 5.11: Reactions of Acids and Bases unit conversions
2.1: Indivisible - The Atomic Theory neutron 1.19: Problem Solving and Unit Conversions
density 2.3: The Properties of Protons, Neutrons, and units of energy
1.22: Density Electrons
1.12: Energy
V viscosity W
valence shell electron pair repulsion 4.11: Intermolecular Forces in Action- Surface Water
Tension and Viscosity
theory 4.15: Water - A Remarkable Molecule
VSEPR
2.22: Predicting the Shapes of Molecules work
2.22: Predicting the Shapes of Molecules
1.12: Energy
Glossary
Sample Word 1 | Sample Definition 1
Detailed Licensing
Overview
Title: Bellingham Technical College CHEM 121
Webpages: 112
Applicable Restrictions: Noncommercial
All licenses found:
CK-12 License: 65.2% (73 pages)
Undeclared: 16.1% (18 pages)
CC BY-NC-SA 3.0: 12.5% (14 pages)
Unknown License: 2.7% (3 pages)
CC BY-SA 4.0: 1.8% (2 pages)
CC BY-SA 3.0: 0.9% (1 page)
Public Domain: 0.9% (1 page)
Unknown License: 0% (0 page)
Unknown License: 0% (0 page)

By Page
Bellingham Technical College CHEM 121 - Undeclared 1.15: Scientific Notation - Writing Large and Small
Front Matter - Undeclared Numbers - CK-12 License
TitlePage - Undeclared 1.16: Significant Figures - Writing Numbers to
InfoPage - Undeclared Reflect Precision - CK-12 License
Table of Contents - Undeclared 1.17: Significant Figures in Calculations - CK-12
Licensing - Undeclared License
1.18: The Basic Units of Measurement - CK-12
1: Module 1 - Matter and Measurement - Undeclared
License
1.1: Chemicals Compose Ordinary Things - CK-12 1.19: Problem Solving and Unit Conversions - CC
License BY-NC-SA 3.0
1.2: The Scope of Chemistry - CK-12 License 1.20: Solving Multi-step Conversion Problems - CC
1.3: What is Matter? - CK-12 License BY-NC-SA 3.0
1.4: Classifying Matter According to Its State—Solid, 1.21: Units Raised to a Power - CK-12 License
Liquid, and Gas - CK-12 License 1.22: Density - CK-12 License
1.5: Classifying Matter According to Its State—Solid,
2: Module 2 - Atoms and Bonding - Undeclared
Liquid, and Gas - CK-12 License
2.1: Indivisible - The Atomic Theory - CK-12 License
1.6: Classifying Matter According to Its Composition
2.2: The Nuclear Atom - CK-12 License
- CK-12 License
2.3: The Properties of Protons, Neutrons, and
1.7: Differences in Matter- Physical and Chemical
Electrons - CK-12 License
Properties - CK-12 License
2.4: Elements- Defined by Their Number of Protons -
1.8: Changes in Matter - Physical and Chemical
CK-12 License
Changes - CK-12 License
2.5: Elements- Defined by Their Number of Protons -
1.9: Conservation of Mass - There is No New Matter
CK-12 License
- CK-12 License
2.6: Looking for Patterns - The Periodic Table - CK-
1.10: Compounds Display Constant Composition -
12 License
CK-12 License
2.7: Ions - Losing and Gaining Electrons - CK-12
1.11: Chemical Formulas - How to Represent
License
Compounds - CK-12 License
2.8: Isotopes - When the Number of Neutrons Varies -
1.12: Energy - CK-12 License
CK-12 License
1.13: Energy and Chemical and Physical Change -
2.9: Atomic Mass - The Average Mass of an
CK-12 License
Element’s Atoms - CK-12 License
1.14: Taking Measurements - CC BY-NC-SA 3.0

1 https://chem.libretexts.org/@go/page/423711
2.10: Light- Electromagnetic Radiation - CK-12 3.9: Making Pancakes- Relationships Between
License Ingredients - Undeclared
2.11: The Electromagnetic Spectrum - CC BY-NC-SA 3.10: Making Molecules- Mole-to-Mole Conversions
3.0 - CC BY-NC-SA 3.0
2.12: The Bohr Model - Atoms with Orbits - CK-12 3.11: Making Molecules- Mass-to-Mass Conversions
License - CC BY-NC-SA 3.0
2.13: The Quantum-Mechanical Model- Atoms with 3.12: Stoichiometry - CK-12 License
Orbitals - CK-12 License 3.13: Limiting Reactant and Theoretical Yield - CC
2.14: Quantum-Mechanical Orbitals and Electron BY-NC-SA 3.0
Configurations - CK-12 License 3.14: Limiting Reactant, Theoretical Yield, and
2.15: The Explanatory Power of the Quantum- Percent Yield from Initial Masses of Reactants - CC
Mechanical Model - CK-12 License BY-NC-SA 3.0
2.16: Periodic Trends - Atomic Size, Ionization 3.15: Enthalpy Change is a Measure of the Heat
Energy, and Metallic Character - CK-12 License Evolved or Absorbed - CK-12 License
2.17: Electron Configurations and the Periodic Table 4: Module 4 - Gases and Liquids - Undeclared
- CK-12 License 4.1: Kinetic Molecular Theory- A Model for Gases -
2.18: Representing Valence Electrons with Dots - CC CC BY-NC-SA 3.0
BY-NC-SA 3.0 4.2: Pressure - The Result of Constant Molecular
2.19: Lewis Structures of Ionic Compounds- Collisions - CK-12 License
Electrons Transferred - CC BY-NC-SA 3.0 4.3: Boyle’s Law - Pressure and Volume - CK-12
2.20: Covalent Lewis Structures- Electrons Shared - License
Unknown License 4.4: Charles’s Law- Volume and Temperature - CK-12
2.21: Writing Lewis Structures for Covalent License
Compounds - Unknown License 4.5: Gay-Lussac's Law- Temperature and Pressure -
2.22: Predicting the Shapes of Molecules - Unknown CK-12 License
License 4.6: Avogadro’s Law- Volume and Moles - CK-12
2.23: Electronegativity and Polarity - Why Oil and License
Water Do not Mix - CK-12 License 4.7: The Ideal Gas Law- Pressure, Volume,
2.24: A Molecular View of Elements and Compounds Temperature, and Moles - CK-12 License
- CK-12 License 4.8: Mixtures of Gases - Why Deep-Sea Divers
2.25: Writing Formulas for Ionic Compounds - CK-12 Breathe a Mixture of Helium and Oxygen - CK-12
License License
2.26: Nomenclature- Naming Compounds - CK-12 4.9: Interactions between Molecules - CK-12 License
License 4.10: Properties of Liquids and Solids - CK-12
2.27: Naming Ionic Compounds - CK-12 License License
2.28: Naming Molecular Compounds - CK-12 4.11: Intermolecular Forces in Action- Surface
License Tension and Viscosity - CK-12 License
3: Module 3 - Stoichiometry - Undeclared 4.12: Evaporation and Condensation - CK-12 License
3.1: Counting Nails by the Pound - CC BY-SA 3.0 4.13: Melting, Freezing, and Sublimation - CK-12
3.2: Counting Atoms by the Gram - CK-12 License License
3.3: Counting Molecules by the Gram - CC BY-NC- 4.14: Intermolecular Forces- Dispersion, Dipole–
SA 3.0 Dipole, Hydrogen Bonding, and Ion-Dipole - CC BY-
3.4: Chemical Formulas as Conversion Factors - CK- NC-SA 3.0
12 License 4.15: Water - A Remarkable Molecule - CK-12
3.5: Mass Percent Composition of Compounds - CK- License
12 License 5: Module 5 - Solution Chemistry - Undeclared
3.6: Mass Percent Composition from a Chemical
5.1: Solutions - Homogeneous Mixtures - CC BY-NC-
Formula - CK-12 License
SA 3.0
3.7: The Chemical Equation - CK-12 License
5.2: Solutions of Solids Dissolved in Water- How to
3.8: How to Write Balanced Chemical Equations -
Make Rock Candy - CK-12 License
CK-12 License
5.3: Specifying Solution Concentration- Mass Percent
- CK-12 License

2 https://chem.libretexts.org/@go/page/423711
5.4: Specifying Solution Concentration- Molarity - 5.13: Acid–Base Titration - CK-12 License
CK-12 License 5.14: Water - Acid and Base in One - CK-12 License
5.5: Solution Dilution - CC BY-SA 4.0 5.15: The pH and pOH Scales - Ways to Express
5.6: Solution Stoichiometry - CC BY-SA 4.0 Acidity and Basicity - CK-12 License
5.7: Electrolytes - Undeclared 5.16: Buffers- Solutions that Resist pH Change - CK-
5.8: Acids- Properties and Examples - CK-12 License 12 License
5.9: Bases- Properties and Examples - CK-12 License Back Matter - Undeclared
5.10: Molecular Definitions of Acids and Bases - CK- Index - Undeclared
12 License Glossary - Undeclared
5.11: Reactions of Acids and Bases - Public Domain Detailed Licensing - Undeclared
5.12: Strong and Weak Acids and Bases - CK-12 Detailed Licensing - Undeclared
License

3 https://chem.libretexts.org/@go/page/423711
Detailed Licensing
Overview
Title: Bellingham Technical College CHEM 121
Webpages: 112
Applicable Restrictions: Noncommercial
All licenses found:
CK-12 License: 65.2% (73 pages)
Undeclared: 16.1% (18 pages)
CC BY-NC-SA 3.0: 12.5% (14 pages)
Unknown License: 2.7% (3 pages)
CC BY-SA 4.0: 1.8% (2 pages)
CC BY-SA 3.0: 0.9% (1 page)
Public Domain: 0.9% (1 page)
Unknown License: 0% (0 page)
Unknown License: 0% (0 page)

By Page
Bellingham Technical College CHEM 121 - Undeclared 1.15: Scientific Notation - Writing Large and Small
Front Matter - Undeclared Numbers - CK-12 License
TitlePage - Undeclared 1.16: Significant Figures - Writing Numbers to
InfoPage - Undeclared Reflect Precision - CK-12 License
Table of Contents - Undeclared 1.17: Significant Figures in Calculations - CK-12
Licensing - Undeclared License
1.18: The Basic Units of Measurement - CK-12
1: Module 1 - Matter and Measurement - Undeclared
License
1.1: Chemicals Compose Ordinary Things - CK-12 1.19: Problem Solving and Unit Conversions - CC
License BY-NC-SA 3.0
1.2: The Scope of Chemistry - CK-12 License 1.20: Solving Multi-step Conversion Problems - CC
1.3: What is Matter? - CK-12 License BY-NC-SA 3.0
1.4: Classifying Matter According to Its State—Solid, 1.21: Units Raised to a Power - CK-12 License
Liquid, and Gas - CK-12 License 1.22: Density - CK-12 License
1.5: Classifying Matter According to Its State—Solid,
2: Module 2 - Atoms and Bonding - Undeclared
Liquid, and Gas - CK-12 License
2.1: Indivisible - The Atomic Theory - CK-12 License
1.6: Classifying Matter According to Its Composition
2.2: The Nuclear Atom - CK-12 License
- CK-12 License
2.3: The Properties of Protons, Neutrons, and
1.7: Differences in Matter- Physical and Chemical
Electrons - CK-12 License
Properties - CK-12 License
2.4: Elements- Defined by Their Number of Protons -
1.8: Changes in Matter - Physical and Chemical
CK-12 License
Changes - CK-12 License
2.5: Elements- Defined by Their Number of Protons -
1.9: Conservation of Mass - There is No New Matter
CK-12 License
- CK-12 License
2.6: Looking for Patterns - The Periodic Table - CK-
1.10: Compounds Display Constant Composition -
12 License
CK-12 License
2.7: Ions - Losing and Gaining Electrons - CK-12
1.11: Chemical Formulas - How to Represent
License
Compounds - CK-12 License
2.8: Isotopes - When the Number of Neutrons Varies -
1.12: Energy - CK-12 License
CK-12 License
1.13: Energy and Chemical and Physical Change -
2.9: Atomic Mass - The Average Mass of an
CK-12 License
Element’s Atoms - CK-12 License
1.14: Taking Measurements - CC BY-NC-SA 3.0

1 https://chem.libretexts.org/@go/page/423714
2.10: Light- Electromagnetic Radiation - CK-12 3.9: Making Pancakes- Relationships Between
License Ingredients - Undeclared
2.11: The Electromagnetic Spectrum - CC BY-NC-SA 3.10: Making Molecules- Mole-to-Mole Conversions
3.0 - CC BY-NC-SA 3.0
2.12: The Bohr Model - Atoms with Orbits - CK-12 3.11: Making Molecules- Mass-to-Mass Conversions
License - CC BY-NC-SA 3.0
2.13: The Quantum-Mechanical Model- Atoms with 3.12: Stoichiometry - CK-12 License
Orbitals - CK-12 License 3.13: Limiting Reactant and Theoretical Yield - CC
2.14: Quantum-Mechanical Orbitals and Electron BY-NC-SA 3.0
Configurations - CK-12 License 3.14: Limiting Reactant, Theoretical Yield, and
2.15: The Explanatory Power of the Quantum- Percent Yield from Initial Masses of Reactants - CC
Mechanical Model - CK-12 License BY-NC-SA 3.0
2.16: Periodic Trends - Atomic Size, Ionization 3.15: Enthalpy Change is a Measure of the Heat
Energy, and Metallic Character - CK-12 License Evolved or Absorbed - CK-12 License
2.17: Electron Configurations and the Periodic Table 4: Module 4 - Gases and Liquids - Undeclared
- CK-12 License 4.1: Kinetic Molecular Theory- A Model for Gases -
2.18: Representing Valence Electrons with Dots - CC CC BY-NC-SA 3.0
BY-NC-SA 3.0 4.2: Pressure - The Result of Constant Molecular
2.19: Lewis Structures of Ionic Compounds- Collisions - CK-12 License
Electrons Transferred - CC BY-NC-SA 3.0 4.3: Boyle’s Law - Pressure and Volume - CK-12
2.20: Covalent Lewis Structures- Electrons Shared - License
Unknown License 4.4: Charles’s Law- Volume and Temperature - CK-12
2.21: Writing Lewis Structures for Covalent License
Compounds - Unknown License 4.5: Gay-Lussac's Law- Temperature and Pressure -
2.22: Predicting the Shapes of Molecules - Unknown CK-12 License
License 4.6: Avogadro’s Law- Volume and Moles - CK-12
2.23: Electronegativity and Polarity - Why Oil and License
Water Do not Mix - CK-12 License 4.7: The Ideal Gas Law- Pressure, Volume,
2.24: A Molecular View of Elements and Compounds Temperature, and Moles - CK-12 License
- CK-12 License 4.8: Mixtures of Gases - Why Deep-Sea Divers
2.25: Writing Formulas for Ionic Compounds - CK-12 Breathe a Mixture of Helium and Oxygen - CK-12
License License
2.26: Nomenclature- Naming Compounds - CK-12 4.9: Interactions between Molecules - CK-12 License
License 4.10: Properties of Liquids and Solids - CK-12
2.27: Naming Ionic Compounds - CK-12 License License
2.28: Naming Molecular Compounds - CK-12 4.11: Intermolecular Forces in Action- Surface
License Tension and Viscosity - CK-12 License
3: Module 3 - Stoichiometry - Undeclared 4.12: Evaporation and Condensation - CK-12 License
3.1: Counting Nails by the Pound - CC BY-SA 3.0 4.13: Melting, Freezing, and Sublimation - CK-12
3.2: Counting Atoms by the Gram - CK-12 License License
3.3: Counting Molecules by the Gram - CC BY-NC- 4.14: Intermolecular Forces- Dispersion, Dipole–
SA 3.0 Dipole, Hydrogen Bonding, and Ion-Dipole - CC BY-
3.4: Chemical Formulas as Conversion Factors - CK- NC-SA 3.0
12 License 4.15: Water - A Remarkable Molecule - CK-12
3.5: Mass Percent Composition of Compounds - CK- License
12 License 5: Module 5 - Solution Chemistry - Undeclared
3.6: Mass Percent Composition from a Chemical
5.1: Solutions - Homogeneous Mixtures - CC BY-NC-
Formula - CK-12 License
SA 3.0
3.7: The Chemical Equation - CK-12 License
5.2: Solutions of Solids Dissolved in Water- How to
3.8: How to Write Balanced Chemical Equations -
Make Rock Candy - CK-12 License
CK-12 License
5.3: Specifying Solution Concentration- Mass Percent
- CK-12 License

2 https://chem.libretexts.org/@go/page/423714
5.4: Specifying Solution Concentration- Molarity - 5.13: Acid–Base Titration - CK-12 License
CK-12 License 5.14: Water - Acid and Base in One - CK-12 License
5.5: Solution Dilution - CC BY-SA 4.0 5.15: The pH and pOH Scales - Ways to Express
5.6: Solution Stoichiometry - CC BY-SA 4.0 Acidity and Basicity - CK-12 License
5.7: Electrolytes - Undeclared 5.16: Buffers- Solutions that Resist pH Change - CK-
5.8: Acids- Properties and Examples - CK-12 License 12 License
5.9: Bases- Properties and Examples - CK-12 License Back Matter - Undeclared
5.10: Molecular Definitions of Acids and Bases - CK- Index - Undeclared
12 License Glossary - Undeclared
5.11: Reactions of Acids and Bases - Public Domain Detailed Licensing - Undeclared
5.12: Strong and Weak Acids and Bases - CK-12 Detailed Licensing - Undeclared
License

3 https://chem.libretexts.org/@go/page/423714

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy