Full
Full
GENERAL CHEMISTRY I
OT - PDX - Metro: General Chemistry I
This text is disseminated via the Open Education Resource (OER) LibreTexts Project (https://LibreTexts.org) and like the hundreds
of other texts available within this powerful platform, it is freely available for reading, printing and "consuming." Most, but not all,
pages in the library have licenses that may allow individuals to make changes, save, and print this book. Carefully
consult the applicable license(s) before pursuing such effects.
Instructors can adopt existing LibreTexts texts or Remix them to quickly build course-specific resources to meet the needs of their
students. Unlike traditional textbooks, LibreTexts’ web based origins allow powerful integration of advanced features and new
technologies to support learning.
The LibreTexts mission is to unite students, faculty and scholars in a cooperative effort to develop an easy-to-use online platform
for the construction, customization, and dissemination of OER content to reduce the burdens of unreasonable textbook costs to our
students and society. The LibreTexts project is a multi-institutional collaborative venture to develop the next generation of open-
access texts to improve postsecondary education at all levels of higher learning by developing an Open Access Resource
environment. The project currently consists of 14 independently operating and interconnected libraries that are constantly being
optimized by students, faculty, and outside experts to supplant conventional paper-based books. These free textbook alternatives are
organized within a central environment that is both vertically (from advance to basic level) and horizontally (across different fields)
integrated.
The LibreTexts libraries are Powered by MindTouch® and are supported by the Department of Education Open Textbook Pilot
Project, the UC Davis Office of the Provost, the UC Davis Library, the California State University Affordable Learning Solutions
Program, and Merlot. This material is based upon work supported by the National Science Foundation under Grant No. 1246120,
1525057, and 1413739. Unless otherwise noted, LibreTexts content is licensed by CC BY-NC-SA 3.0.
Any opinions, findings, and conclusions or recommendations expressed in this material are those of the author(s) and do not
necessarily reflect the views of the National Science Foundation nor the US Department of Education.
Have questions or comments? For information about adoptions or adaptions contact [email protected]. More information on our
activities can be found via Facebook (https://facebook.com/Libretexts), Twitter (https://twitter.com/libretexts), or our blog
(http://Blog.Libretexts.org).
1 https://chem.libretexts.org/@go/page/254521
5: Transformations of Matter
5.1: Writing and Balancing Chemical Equations
5.1.1: Practice Problems- Writing and Balancing Chemical Equations
5.2: Reaction Stoichiometry
5.2.1: Practice Problems- Reaction Stoichiometry
5.3: Calculating Reaction Yields
5.3.1: Practice Problems Calculating Reaction Yields
6: Aqueous Reactions
6.1: Solutions and Solution Concentration
6.1.1: Practice Problems- Solution Concentration
6.2: Solutions Chemistry
6.2.1: Practice Problems- Solutions Chemistry
6.3: Precipitation Reactions
6.3.1: Practice Problems- Precipitation
6.4: Classifying Chemical Reactions (Acids and Bases)
6.4.1: Classifying Chemical Reactions (Acids and Bases) (Problems)
6.5: Classifying Chemical Reactions (Redox)
6.5.1: Classifying Chemical Reactions (Redox) (Problems)
8: Thermochemistry
8.1: The Basics of Energy
8.1.1: Practice Problems- The Basics of Energy
8.2: Measuring Energy and Heat Capacity
8.3: Enthalpy
8.4: Standard Enthalpy and Hess’ Law
8.4.1: Practice Problems- Enthalpy and Hess’ Law
8.5: Calorimetry
8.5.1: Practice Problems- Calorimetry
Index
Glossary
2 https://chem.libretexts.org/@go/page/254521
CHAPTER OVERVIEW
1: Matter and Measurement
1.1: Atoms and Molecules
1.2: The Scientific Approach to Knowledge
1.3: The Classification of Matter
1.4: Measurements
1.4.1: Practice Problems on Measurements
1.5: The Mole is a Measure of Amount
1.6: Accuracy and Precision
1.7: Dimensional Analysis
1.7.1: Practice Problems on Dimensional Analysis
1.8: Significant Digits
1: Matter and Measurement is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
1
1.1: Atoms and Molecules
Skills to Develop
Outline the historical development of chemistry
Provide examples of the importance of chemistry in everyday life
Provide examples illustrating macroscopic, microscopic, and symbolic domains
Throughout human history, people have tried to convert matter into more useful forms. Our Stone Age ancestors chipped pieces of
flint into useful tools and carved wood into statues and toys. These endeavors involved changing the shape of a substance without
changing the substance itself. But as our knowledge increased, humans began to change the composition of the substances as well
—clay was converted into pottery, hides were cured to make garments, copper ores were transformed into copper tools and
weapons, and grain was made into bread.
Humans began to practice chemistry when they learned to control fire and use it to cook, make pottery, and smelt metals.
Subsequently, they began to separate and use specific components of matter. A variety of drugs such as aloe, myrrh, and opium
were isolated from plants. Dyes, such as indigo and Tyrian purple, were extracted from plant and animal matter. Metals were
combined to form alloys—for example, copper and tin were mixed together to make bronze—and more elaborate smelting
techniques produced iron. Alkalis were extracted from ashes, and soaps were prepared by combining these alkalis with fats.
Alcohol was produced by fermentation and purified by distillation.
Attempts to understand the behavior of matter extend back for more than 2500 years. As early as the sixth century BC, Greek
philosophers discussed a system in which water was the basis of all things. You may have heard of the Greek postulate that matter
consists of four elements: earth, air, fire, and water. Subsequently, an amalgamation of chemical technologies and philosophical
speculations were spread from Egypt, China, and the eastern Mediterranean by alchemists, who endeavored to transform “base
metals” such as lead into “noble metals” like gold, and to create elixirs to cure disease and extend life (Figure 1.1.1).
Figure 1.1.1: This portrayal shows an alchemist’s workshop circa 1580. Although alchemy made some useful contributions to how
to manipulate matter, it was not scientific by modern standards. (credit: Chemical Heritage Foundation).
From alchemy came the historical progressions that led to modern chemistry: the isolation of drugs from natural sources,
metallurgy, and the dye industry. Today, chemistry continues to deepen our understanding and improve our ability to harness and
control the behavior of matter. This effort has been so successful that many people do not realize either the central position of
chemistry among the sciences or the importance and universality of chemistry in daily life.
between many subdisciplines within the two fields, such as chemical physics and nuclear chemistry. Mathematics, computer
science, and information theory provide important tools that help us calculate, interpret, describe, and generally make sense of the
chemical world. Biology and chemistry converge in biochemistry, which is crucial to understanding the many complex factors and
processes that keep living organisms (such as us) alive. Chemical engineering, materials science, and nanotechnology combine
chemical principles and empirical findings to produce useful substances, ranging from gasoline to fabrics to electronics.
Agriculture, food science, veterinary science, and brewing and wine making help provide sustenance in the form of food and drink
1.1.1 https://chem.libretexts.org/@go/page/218243
to the world’s population. Medicine, pharmacology, biotechnology, and botany identify and produce substances that help keep us
healthy. Environmental science, geology, oceanography, and atmospheric science incorporate many chemical ideas to help us better
understand and protect our physical world. Chemical ideas are used to help understand the universe in astronomy and cosmology.
Figure 1.1.2: Knowledge of chemistry is central to understanding a wide range of scientific disciplines. This diagram shows just
some of the interrelationships between chemistry and other fields.
What are some changes in matter that are essential to daily life? Digesting and assimilating food, synthesizing polymers that are
used to make clothing, containers, cookware, and credit cards, and refining crude oil into gasoline and other products are just a few
examples. As you proceed through this course, you will discover many different examples of changes in the composition and
structure of matter, how to classify these changes and how they occurred, their causes, the changes in energy that accompany them,
and the principles and laws involved. As you learn about these things, you will be learning chemistry, the study of the composition,
properties, and interactions of matter. The practice of chemistry is not limited to chemistry books or laboratories: It happens
whenever someone is involved in changes in matter or in conditions that may lead to such changes.
1.1.2 https://chem.libretexts.org/@go/page/218243
A helpful way to understand the three domains is via the essential and ubiquitous substance of water. That water is a liquid at
moderate temperatures, will freeze to form a solid at lower temperatures, and boil to form a gas at higher temperatures (Figure
1.1.3) are macroscopic observations. But some properties of water fall into the microscopic domain—what we cannot observe with
the naked eye. The description of water as comprised of two hydrogen atoms and one oxygen atom, and the explanation of freezing
and boiling in terms of attractions between these molecules, is within the microscopic arena. The formula H2O, which can describe
water at either the macroscopic or microscopic levels, is an example of the symbolic domain. The abbreviations (g) for gas, (s) for
solid, and (l) for liquid are also symbolic.
Figure 1.1.3: (a) Moisture in the air, icebergs, and the ocean represent water in the macroscopic domain. (b) At the molecular level
(microscopic domain), gas molecules are far apart and disorganized, solid water molecules are close together and organized, and
liquid molecules are close together and disorganized. (c) The formula H2O symbolizes water, and (g), (s), and (l) symbolize its
phases. Note that clouds are actually comprised of either very small liquid water droplets or solid water crystals; gaseous water in
our atmosphere is not visible to the naked eye, although it may be sensed as humidity. (credit a: modification of work by
“Gorkaazk”/Wikimedia Commons).
Summary
Chemistry deals with the composition, structure, and properties of matter, and the ways by which various forms of matter may be
interconverted. Thus, it occupies a central place in the study and practice of science and technology. Chemists use the scientific
method to perform experiments, pose hypotheses, and formulate laws and develop theories, so that they can better understand the
behavior of the natural world. To do so, they operate in the macroscopic, microscopic, and symbolic domains. Chemists measure,
analyze, purify, and synthesize a wide variety of substances that are important to our lives.
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/[email protected]).
1.1: Atoms and Molecules is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
1.1.3 https://chem.libretexts.org/@go/page/218243
1.2: The Scientific Approach to Knowledge
Skills to Develop
Describe the process of the scientific method
Scientists search for answers to questions and solutions to problems by using a procedure called the scientific method. This
procedure consists of making observations, formulating hypotheses, and designing experiments, which in turn lead to additional
observations, hypotheses, and experiments in repeated cycles (Figure 1.2.1).
Figure 1.2.1 : The Scientific Method. As depicted in this flowchart, the scientific method consists of making observations,
formulating hypotheses, and designing experiments. A scientist may enter the cycle at any point.
Observations can be qualitative or quantitative. Qualitative observations describe properties or occurrences in ways that do not rely
on numbers. Examples of qualitative observations include the following: the outside air temperature is cooler during the winter
season, table salt is a crystalline solid, sulfur crystals are yellow, and dissolving a penny in dilute nitric acid forms a blue solution
and a brown gas. Quantitative observations are measurements, which by definition consist of both a number and a unit. Examples
of quantitative observations include the following: the melting point of crystalline sulfur is 115.21 °C, and 35.9 grams of table salt
—whose chemical name is sodium chloride—dissolve in 100 grams of water at 20 °C. An example of a quantitative observation
was the initial observation leading to the modern theory of the dinosaurs’ extinction: iridium concentrations in sediments dating to
66 million years ago were found to be 20–160 times higher than normal. The development of this theory is a good exemplar of the
scientific method in action (see Figure 1.2.2 below).
After deciding to learn more about an observation or a set of observations, scientists generally begin an investigation by forming a
hypothesis, a tentative explanation for the observation(s). The hypothesis may not be correct, but it puts the scientist’s
understanding of the system being studied into a form that can be tested. For example, the observation that we experience
alternating periods of light and darkness corresponding to observed movements of the sun, moon, clouds, and shadows is consistent
with either of two hypotheses:
1. Earth rotates on its axis every 24 hours, alternately exposing one side to the sun, or
2. The sun revolves around Earth every 24 hours.
Suitable experiments can be designed to choose between these two alternatives. For the disappearance of the dinosaurs, the
hypothesis was that the impact of a large extraterrestrial object caused their extinction. Unfortunately (or perhaps fortunately), this
hypothesis does not lend itself to direct testing by any obvious experiment, but scientists collected additional data that either
support or refute it.
After a hypothesis has been formed, scientists conduct experiments to test its validity. Experiments are systematic observations or
measurements, preferably made under controlled conditions—that is, under conditions in which a single variable changes. For
example, in the dinosaur extinction scenario, iridium concentrations were measured worldwide and compared. A properly designed
and executed experiment enables a scientist to determine whether the original hypothesis is valid. Experiments often demonstrate
that the hypothesis is incorrect or that it must be modified. More experimental data are then collected and analyzed, at which point
1.2.1 https://chem.libretexts.org/@go/page/218244
a scientist may begin to think that the results are sufficiently reproducible (i.e., dependable) to merit being summarized in a law, a
verbal or mathematical description of a phenomenon that allows for general predictions. A law simply says what happens; it does
not address the question of why.
One example of a law, the Law of Definite Proportions, which was discovered by the French scientist Joseph Proust (1754–1826),
states that a chemical substance always contains the same proportions of elements by mass. Thus sodium chloride (table salt)
always contains the same proportion by mass of sodium to chlorine, in this case 39.34% sodium and 60.66% chlorine by mass, and
sucrose (table sugar) is always 42.11% carbon, 6.48% hydrogen, and 51.41% oxygen by mass. Some solid compounds do not
strictly obey the law of definite proportions. The law of definite proportions should seem obvious—we would expect the
composition of sodium chloride to be consistent—but the head of the US Patent Office did not accept it as a fact until the early 20th
century.
Whereas a law states only what happens, a theory attempts to explain why nature behaves as it does. Laws are unlikely to change
greatly over time unless a major experimental error is discovered. In contrast, a theory, by definition, is incomplete and imperfect,
evolving with time to explain new facts as they are discovered. The theory developed to explain the extinction of the dinosaurs, for
example, is that Earth occasionally encounters small- to medium-sized asteroids, and these encounters may have unfortunate
implications for the continued existence of most species. This theory is by no means proven, but it is consistent with the bulk of
evidence amassed to date. Figure 1.2.2 summarizes the application of the scientific method in this case.
Figure 1.2.2 : A Summary of How the Scientific Method Was Used in Developing the Asteroid Impact Theory to Explain the
Disappearance of the Dinosaurs from Earth
Example 1.2.1
Classify each statement as a law, a theory, an experiment, a hypothesis, a qualitative observation, or a quantitative observation.
1.2.2 https://chem.libretexts.org/@go/page/218244
a. Ice always floats on liquid water.
b. Birds evolved from dinosaurs.
c. Hot air is less dense than cold air, probably because the components of hot air are moving more rapidly.
d. When 10 g of ice were added to 100 mL of water at 25 °C, the temperature of the water decreased to 15.5 °C after the ice
melted.
e. The ingredients of Ivory soap were analyzed to see whether it really is 99.44% pure, as advertised.
Given: components of the scientific method
Asked for: statement classification
Strategy: Refer to the definitions in this section to determine which category best describes each statement.
Solution
a. This is a general statement of a relationship between the properties of liquid and solid water, so it is a law.
b. This is a possible explanation for the origin of birds, so it is a hypothesis.
c. This is a statement that tries to explain the relationship between the temperature and the density of air based on
fundamental principles, so it is a theory.
d. The temperature is measured before and after a change is made in a system, so these are quantitative observations.
e. This is an analysis designed to test a hypothesis (in this case, the manufacturer’s claim of purity), so it is an experiment.
Exercise 1.2.1
Classify each statement as a law, a theory, an experiment, a hypothesis, a qualitative observation, or a quantitative observation.
a. Measured amounts of acid were added to a Rolaids tablet to see whether it really “consumes 47 times its weight in excess
stomach acid.”
b. Heat always flows from hot objects to cooler ones, not in the opposite direction.
c. The universe was formed by a massive explosion that propelled matter into a vacuum.
d. Michael Jordan is the greatest pure shooter ever to play professional basketball.
e. Limestone is relatively insoluble in water but dissolves readily in dilute acid with the evolution of a gas.
f. Gas mixtures that contain more than 4% hydrogen in air are potentially explosive.
Answer a
experiment
Answer b
law
Answer c
theory
Answer d
hypothesis
Answer e
qualitative observation
Answer f
quantitative observation
Because scientists can enter the cycle shown in Figure 1.2.1 at any point, the actual application of the scientific method to different
topics can take many different forms. For example, a scientist may start with a hypothesis formed by reading about work done by
others in the field, rather than by making direct observations.
It is important to remember that scientists have a tendency to formulate hypotheses in familiar terms simply because it is difficult to
propose something that has never been encountered or imagined before. As a result, scientists sometimes discount or overlook
unexpected findings that disagree with the basic assumptions behind the hypothesis or theory being tested. Fortunately, truly
1.2.3 https://chem.libretexts.org/@go/page/218244
important findings are immediately subject to independent verification by scientists in other laboratories, so science is a self-
correcting discipline. When the Alvarezes originally suggested that an extraterrestrial impact caused the extinction of the dinosaurs,
the response was almost universal skepticism and scorn. In only 20 years, however, the persuasive nature of the evidence overcame
the skepticism of many scientists, and their initial hypothesis has now evolved into a theory that has revolutionized paleontology
and geology.
Summary
Chemists expand their knowledge by making observations, carrying out experiments, and testing hypotheses to develop laws to
summarize their results and theories to explain them. In doing so, they are using the scientific method.
Contributors
The Scientific Meth. (n.d.). Retrieved June 1, 2020, from https://chem.libretexts.org/Bookshel...ientific_Metho
1.2: The Scientific Approach to Knowledge is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
1.2.4 https://chem.libretexts.org/@go/page/218244
1.3: The Classification of Matter
Skills to Develop
Describe the fundamental properties of each physical state of matter: solid, liquid, and gas
Define and give examples of atoms and molecules
Classify matter as an element, or compound; pure substance, or a mixture; a homogenous mixture, or heterogenous mixture
Classify matter as chemical or physical
Classify properties of matter as chemical, or physical; an extensive or intensive
Distinguish between mass and weight
Matter is defined as anything that occupies space and has mass, and it is all around us. Solids and liquids are more obviously
matter: We can see that they take up space, and their weight tells us that they have mass. Gases are also matter; if gases did not take
up space, a balloon would stay collapsed rather than inflate when filled with gas.
Solids, liquids, and gases are the three states of matter commonly found on earth (Figure 1.3.1). A solid is rigid and possesses a
definite shape. A liquid flows and takes the shape of a container, except that it forms a flat or slightly curved upper surface when
acted upon by gravity. (In zero gravity, liquids assume a spherical shape.) Both liquid and solid samples have volumes that are very
nearly independent of pressure. A gas takes both the shape and volume of its container.
Figure 1.3.1 : The three most common states or phases of matter are solid, liquid, and gas.
A fourth state of matter, plasma, occurs naturally in the interiors of stars. A plasma is a gaseous state of matter that contains
appreciable numbers of electrically charged particles (Figure 1.3.2). The presence of these charged particles imparts unique
properties to plasmas that justify their classification as a state of matter distinct from gases. In addition to stars, plasmas are found
in some other high-temperature environments (both natural and man-made), such as lightning strikes, certain television screens,
and specialized analytical instruments used to detect trace amounts of metals.
Figure 1.3.2 : A plasma torch can be used to cut metal. (credit: “Hypertherm”/Wikimedia Commons)
1.3.1 https://chem.libretexts.org/@go/page/218245
What is Plasma?
Video 1.3.1 : In a tiny cell in a plasma television, the plasma emits ultraviolet light, which in turn causes the display at that
location to appear a specific color. The composite of these tiny dots of color makes up the image that you see. Watch this video to
learn more about plasma and the places you encounter it.
Some samples of matter appear to have properties of solids, liquids, and/or gases at the same time. This can occur when the sample
is composed of many small pieces. For example, we can pour sand as if it were a liquid because it is composed of many small
grains of solid sand. Matter can also have properties of more than one state when it is a mixture, such as with clouds. Clouds appear
to behave somewhat like gases, but they are actually mixtures of air (gas) and tiny particles of water (liquid or solid).
The mass of an object is a measure of the amount of matter in it. One way to measure an object’s mass is to measure the force it
takes to accelerate the object. It takes much more force to accelerate a car than a bicycle because the car has much more mass. A
more common way to determine the mass of an object is to use a balance to compare its mass with a standard mass.
Although weight is related to mass, it is not the same thing. Weight refers to the force that gravity exerts on an object. This force is
directly proportional to the mass of the object. The weight of an object changes as the force of gravity changes, but its mass does
not. An astronaut’s mass does not change just because she goes to the moon. But her weight on the moon is only one-sixth her
earth-bound weight because the moon’s gravity is only one-sixth that of the earth’s. She may feel “weightless” during her trip when
she experiences negligible external forces (gravitational or any other), although she is, of course, never “massless.”
The law of conservation of matter summarizes many scientific observations about matter: It states that there is no detectable
change in the total quantity of matter present when matter converts from one type to another (a chemical change) or changes
among solid, liquid, or gaseous states (a physical change). Brewing beer and the operation of batteries provide examples of the
conservation of matter (Figure 1.3.3). During the brewing of beer, the ingredients (water, yeast, grains, malt, hops, and sugar) are
converted into beer (water, alcohol, carbonation, and flavoring substances) with no actual loss of substance. This is most clearly
seen during the bottling process, when glucose turns into ethanol and carbon dioxide, and the total mass of the substances does not
change. This can also be seen in a lead-acid car battery: The original substances (lead, lead oxide, and sulfuric acid), which are
capable of producing electricity, are changed into other substances (lead sulfate and water) that do not produce electricity, with no
change in the actual amount of matter.
1.3.2 https://chem.libretexts.org/@go/page/218245
Figure 1.3.3 : (a) The mass of beer precursor materials is the same as the mass of beer produced: Sugar has become alcohol and
carbonation. (b) The mass of the lead, lead oxide plates, and sulfuric acid that goes into the production of electricity is exactly
equal to the mass of lead sulfate and water that is formed.
Although this conservation law holds true for all conversions of matter, convincing examples are few and far between because,
outside of the controlled conditions in a laboratory, we seldom collect all of the material that is produced during a particular
conversion. For example, when you eat, digest, and assimilate food, all of the matter in the original food is preserved. But because
some of the matter is incorporated into your body, and much is excreted as various types of waste, it is challenging to verify by
measurement.
Figure 1.3.5 : These images provide an increasingly closer view: (a) a cotton boll, (b) a single cotton fiber viewed under an optical
microscope (magnified 40 times), (c) an image of a cotton fiber obtained with an electron microscope (much higher magnification
than with the optical microscope); and (d and e) atomic-level models of the fiber (spheres of different colors represent atoms of
different elements). (credit c: modification of work by “Featheredtar”/Wikimedia Commons)
1.3.3 https://chem.libretexts.org/@go/page/218245
An atom is so light that its mass is also difficult to imagine. A billion lead atoms (1,000,000,000 atoms) weigh about 3 × 10 −13
grams, a mass that is far too light to be weighed on even the world’s most sensitive balances. It would require over
300,000,000,000,000 lead atoms (300 trillion, or 3 × 1014) to be weighed, and they would weigh only 0.0000001 gram.
It is rare to find collections of individual atoms. Only a few elements, such as the gases helium, neon, and argon, consist of a
collection of individual atoms that move about independently of one another. Other elements, such as the gases hydrogen, nitrogen,
oxygen, and chlorine, are composed of units that consist of pairs of atoms (Figure 1.3.6). One form of the element phosphorus
consists of units composed of four phosphorus atoms. The element sulfur exists in various forms, one of which consists of units
composed of eight sulfur atoms. These units are called molecules. A molecule consists of two or more atoms joined by strong
forces called chemical bonds. The atoms in a molecule move around as a unit, much like the cans of soda in a six-pack or a bunch
of keys joined together on a single key ring. A molecule may consist of two or more identical atoms, as in the molecules found in
the elements hydrogen, oxygen, and sulfur, or it may consist of two or more different atoms, as in the molecules found in water.
Each water molecule is a unit that contains two hydrogen atoms and one oxygen atom. Each glucose molecule is a unit that
contains 6 carbon atoms, 12 hydrogen atoms, and 6 oxygen atoms. Like atoms, molecules are incredibly small and light. If an
ordinary glass of water were enlarged to the size of the earth, the water molecules inside it would be about the size of golf balls.
Figure 1.3.6 : The elements hydrogen, oxygen, phosphorus, and sulfur form molecules consisting of two or more atoms of the same
element. The compounds water, carbon dioxide, and glucose consist of combinations of atoms of different elements.
Classifying Matter
We can classify matter into several categories. Two broad categories are mixtures and pure substances. A pure substance has a
constant composition. All specimens of a pure substance have exactly the same makeup and properties. Any sample of sucrose
(table sugar) consists of 42.1% carbon, 6.5% hydrogen, and 51.4% oxygen by mass. Any sample of sucrose also has the same
physical properties, such as melting point, color, and sweetness, regardless of the source from which it is isolated.
We can divide pure substances into two classes: elements and compounds. Pure substances that cannot be broken down into simpler
substances by chemical changes are called elements. Iron, silver, gold, aluminum, sulfur, oxygen, and copper are familiar examples
of the more than 100 known elements, of which about 90 occur naturally on the earth, and two dozen or so have been created in
laboratories.
Pure substances that can be broken down by chemical changes are called compounds. This breakdown may produce either elements
or other compounds, or both. Mercury(II) oxide, an orange, crystalline solid, can be broken down by heat into the elements
mercury and oxygen (Figure 1.3.7). When heated in the absence of air, the compound sucrose is broken down into the element
carbon and the compound water. (The initial stage of this process, when the sugar is turning brown, is known as caramelization—
this is what imparts the characteristic sweet and nutty flavor to caramel apples, caramelized onions, and caramel). Silver(I) chloride
is a white solid that can be broken down into its elements, silver and chlorine, by absorption of light. This property is the basis for
the use of this compound in photographic films and photochromic eyeglasses (those with lenses that darken when exposed to light).
Figure 1.3.7 : (a)The compound mercury(II) oxide, (b)when heated, (c) decomposes into silvery droplets of liquid mercury and
invisible oxygen gas. (credit: modification of work by Paul Flowers)
1.3.4 https://chem.libretexts.org/@go/page/218245
The properties of combined elements are different from those in the free, or uncombined, state. For example, white crystalline
sugar (sucrose) is a compound resulting from the chemical combination of the element carbon, which is a black solid in one of its
uncombined forms, and the two elements hydrogen and oxygen, which are colorless gases when uncombined. Free sodium, an
element that is a soft, shiny, metallic solid, and free chlorine, an element that is a yellow-green gas, combine to form sodium
chloride (table salt), a compound that is a white, crystalline solid.
A mixture is composed of two or more types of matter that can be present in varying amounts and can be separated by physical
changes, such as evaporation (you will learn more about this later). A mixture with a composition that varies from point to point is
called a heterogeneous mixture. Italian dressing is an example of a heterogeneous mixture (Figure 1.3.8a). Its composition can
vary because we can make it from varying amounts of oil, vinegar, and herbs. It is not the same from point to point throughout the
mixture—one drop may be mostly vinegar, whereas a different drop may be mostly oil or herbs because the oil and vinegar
separate and the herbs settle. Other examples of heterogeneous mixtures are chocolate chip cookies (we can see the separate bits of
chocolate, nuts, and cookie dough) and granite (we can see the quartz, mica, feldspar, and more).
A homogeneous mixture, also called a solution, exhibits a uniform composition and appears visually the same throughout. An
example of a solution is a sports drink, consisting of water, sugar, coloring, flavoring, and electrolytes mixed together uniformly
(Figure 1.3.8b). Each drop of a sports drink tastes the same because each drop contains the same amounts of water, sugar, and other
components. Note that the composition of a sports drink can vary—it could be made with somewhat more or less sugar, flavoring,
or other components, and still be a sports drink. Other examples of homogeneous mixtures include air, maple syrup, gasoline, and a
solution of salt in water.
Figure 1.3.8 : (a) Oil and vinegar salad dressing is a heterogeneous mixture because its composition is not uniform throughout. (b)
A commercial sports drink is a homogeneous mixture because its composition is uniform throughout. (credit a “left”: modification
of work by John Mayer; credit a “right”: modification of work by Umberto Salvagnin; credit b “left: modification of work by Jeff
Bedford)
Although there are just over 100 elements, tens of millions of chemical compounds result from different combinations of these
elements. Each compound has a specific composition and possesses definite chemical and physical properties by which we can
distinguish it from all other compounds. And, of course, there are innumerable ways to combine elements and compounds to form
different mixtures. The overall organization of matter and the methods used to separate mixtures are summarized in Figure 1.3.6.
Figure 1.3.6 : Relationships between the Types of Matter and the Methods Used to Separate Mixtures
1.3.5 https://chem.libretexts.org/@go/page/218245
Example 1.3.1
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 1.3.1
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)
Answer A
solution
Answer B
element
Answer C
heterogeneous mixture
Answer D
compound
1.3.6 https://chem.libretexts.org/@go/page/218245
Physical and Chemical Properties of Matter
The characteristics that enable us to distinguish one substance from another are called properties. A physical property is a
characteristic of matter that is not associated with a change in its chemical composition. Familiar examples of physical properties
include density, color, hardness, melting and boiling points, and electrical conductivity. We can observe some physical properties,
such as density and color, without changing the physical state of the matter observed. Other physical properties, such as the melting
temperature of iron or the freezing temperature of water, can only be observed as matter undergoes a physical change. A physical
change is a change in the state or properties of matter without any accompanying change in its chemical composition (the identities
of the substances contained in the matter). We observe a physical change when wax melts, when sugar dissolves in coffee, and
when steam condenses into liquid water (Figure 1.3.1). Other examples of physical changes include magnetizing and
demagnetizing metals (as is done with common antitheft security tags) and grinding solids into powders (which can sometimes
yield noticeable changes in color). In each of these examples, there is a change in the physical state, form, or properties of the
substance, but no change in its chemical composition.
Figure 1.3.1 : (a) Wax undergoes a physical change when solid wax is heated and forms liquid wax. (b) Steam condensing inside a
cooking pot is a physical change, as water vapor is changed into liquid water. (credit a: modification of work by
“95jb14”/Wikimedia Commons; credit b: modification of work by “mjneuby”/Flickr).
The change of one type of matter into another type (or the inability to change) is a chemical property. Examples of chemical
properties include flammability, toxicity, acidity, reactivity (many types), and heat of combustion. Iron, for example, combines with
oxygen in the presence of water to form rust; chromium does not oxidize (Figure 1.3.2). Nitroglycerin is very dangerous because it
explodes easily; neon poses almost no hazard because it is very unreactive.
Figure 1.3.2 : (a) One of the chemical properties of iron is that it rusts; (b) one of the chemical properties of chromium is that it
does not. (credit a: modification of work by Tony Hisgett; credit b: modification of work by “Atoma”/Wikimedia Commons)
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. The explosion of nitroglycerin is a chemical change because
the gases produced are very different kinds of matter from the original substance. Other examples of chemical changes include
reactions that are performed in a lab (such as copper reacting with nitric acid), all forms of combustion (burning), and food being
cooked, digested, or rotting.
Properties of matter fall into one of two categories. If the property depends on the amount of matter present, it is an extensive
property. The mass and volume of a substance are examples of extensive properties; for instance, a gallon of milk has a larger mass
and volume than a cup of milk. The value of an extensive property is directly proportional to the amount of matter in question. If
the property of a sample of matter does not depend on the amount of matter present, it is an intensive property. Temperature is an
example of an intensive property. If the gallon and cup of milk are each at 20 °C (room temperature), when they are combined, the
1.3.7 https://chem.libretexts.org/@go/page/218245
temperature remains at 20 °C. As another example, consider the distinct but related properties of heat and temperature. A drop of
hot cooking oil spattered on your arm causes brief, minor discomfort, whereas a pot of hot oil yields severe burns. Both the drop
and the pot of oil are at the same temperature (an intensive property), but the pot clearly contains much more heat (extensive
property).
While many elements differ dramatically in their chemical and physical properties, some elements have similar properties. We can
identify sets of elements that exhibit common behaviors. For example, many elements conduct heat and electricity well, whereas
others are poor conductors. These properties can be used to sort the elements into three classes: metals (elements that conduct
well), nonmetals (elements that conduct poorly), and metalloids (elements that have properties of both metals and nonmetals).
The periodic table is a table of elements that places elements with similar properties close together (Figure 1.3.5). You will learn
more about the periodic table as you continue your study of chemistry.
Figure 1.3.5 : The periodic table shows how elements may be grouped according to certain similar properties. Note the background
color denotes whether an element is a metal, metalloid, or nonmetal, whereas the element symbol color indicates whether it is a
solid, liquid, or gas.
Glossary
atom
smallest particle of an element that can enter into a chemical combination
chemical change
change producing a different kind of matter from the original kind of matter
chemical property
behavior that is related to the change of one kind of matter into another kind of matter
compound
pure substance that can be decomposed into two or more elements
element
substance that is composed of a single type of atom; a substance that cannot be decomposed by a chemical change
1.3.8 https://chem.libretexts.org/@go/page/218245
extensive property
property of a substance that depends on the amount of the substance
gas
state in which matter has neither definite volume nor shape
heterogeneous mixture
combination of substances with a composition that varies from point to point
homogeneous mixture
(also, solution) combination of substances with a composition that is uniform throughout
intensive property
property of a substance that is independent of the amount of the substance
liquid
state of matter that has a definite volume but indefinite shape
law of conservation of matter
when matter converts from one type to another or changes form, there is no detectable change in the total amount of matter
present
mass
fundamental property indicating amount of matter
matter
anything that occupies space and has mass
mixture
matter that can be separated into its components by physical means
molecule
bonded collection of two or more atoms of the same or different elements
physical change
change in the state or properties of matter that does not involve a change in its chemical composition
physical property
characteristic of matter that is not associated with any change in its chemical composition
plasma
gaseous state of matter containing a large number of electrically charged atoms and/or molecules
pure substance
homogeneous substance that has a constant composition
solid
state of matter that is rigid, has a definite shape, and has a fairly constant volume
weight
force that gravity exerts on an object
Summary
Matter is anything that occupies space and has mass. The basic building block of matter is the atom, the smallest unit of an element
that can enter into combinations with atoms of the same or other elements. In many substances, atoms are combined into
molecules. On earth, matter commonly exists in three states: solids, of fixed shape and volume; liquids, of variable shape but fixed
volume; and gases, of variable shape and volume. Under high-temperature conditions, matter also can exist as a plasma. Most
matter is a mixture: It is composed of two or more types of matter that can be present in varying amounts and can be separated by
physical means. Heterogeneous mixtures vary in composition from point to point; homogeneous mixtures have the same
composition from point to point. Pure substances consist of only one type of matter. A pure substance can be an element, which
consists of only one type of atom and cannot be broken down by a chemical change, or a compound, which consists of two or more
types of atoms.
1.3.9 https://chem.libretexts.org/@go/page/218245
Matter can be classified according to physical and chemical properties. Matter is anything that occupies space and has mass. The
three states of matter are solid, liquid, and gas. A physical change involves the conversion of a substance from one state of matter
to another, without changing its chemical composition. Most matter consists of mixtures of pure substances, which can be
homogeneous (uniform in composition) or heterogeneous (different regions possess different compositions and properties). Pure
substances can be either chemical compounds or elements. Compounds can be broken down into elements by chemical reactions,
but elements cannot be separated into simpler substances by chemical means. The properties of substances can be classified as
either physical or chemical. Scientists can observe physical properties without changing the composition of the substance, whereas
chemical properties describe the tendency of a substance to undergo chemical changes (chemical reactions) that change its
chemical composition. Physical properties can be intensive or extensive. Intensive properties are the same for all samples; do not
depend on sample size; and include, for example, color, physical state, and melting and boiling points. Extensive properties depend
on the amount of material and include mass and volume. The ratio of two extensive properties, mass and volume, is an important
intensive property called density.
1.3: The Classification of Matter is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
1.3.10 https://chem.libretexts.org/@go/page/218245
1.4: Measurements
Skills to Develop
Explain the process of measurement
Identify the two basic parts of a quantity
Describe the properties and units of length, mass, volume, density, and time
Perform basic unit calculations and conversions in the metric and other unit systems
Video 1.4.1 : Watch this video for an introduction to units for this unit
Measurements provide the macroscopic information that is the basis of most of the hypotheses, theories, and laws that describe the
behavior of matter and energy in both the macroscopic and microscopic domains of chemistry. For our purposes, every
measurement provides two kinds of information: the size or magnitude of the measurement (a number or quantity); and a standard
of comparison for the measurement (a unit). The number and unit are explicitly represented when a quantity is written.
The number in the measurement can be represented in different ways, including decimal form and scientific notation. For example,
the maximum takeoff weight of a Boeing 777-200ER airliner is 298,000 kilograms, which can also be written as 2.98 × 105 kg.
The mass of the average mosquito is about 0.0000025 kilograms, which can be written as 2.5 × 10−6 kg.
Units, such as liters, pounds, and centimeters, are standards of comparison for measurements. When we buy a 2-liter bottle of a soft
drink, we expect that the volume of the drink was measured, so it is two times larger than the volume that everyone agrees to be 1
liter. The meat used to prepare a 0.25-pound hamburger is measured so it weighs one-fourth as much as 1 pound. Without units, a
number can be meaningless, confusing, or possibly life threatening. Suppose a doctor prescribes phenobarbital to control a patient’s
seizures and states a dosage of “100” without specifying units. Not only will this be confusing to the medical professional giving
the dose, but the consequences can be dire: 100 mg given three times per day can be effective as an anticonvulsant, but a single
dose of 100 g is more than 10 times the lethal amount.
We usually report the results of scientific measurements in SI units, an updated version of the metric system, using the units listed
in Table 1.4.1. Other units can be derived from these base units. The standards for these units are fixed by international agreement,
and they are called the International System of Units or SI Units (from the French, Le Système International d’Unités). SI units
have been used by the United States National Institute of Standards and Technology (NIST) since 1964.
Table 1.4.1: Base Units of the SI System
Property Measured Name of Unit Symbol of Unit
length meter m
mass kilogram kg
time second s
temperature kelvin K
1.4.1 https://chem.libretexts.org/@go/page/217238
Property Measured Name of Unit Symbol of Unit
Sometimes we use units that are fractions or multiples of a base unit. Ice cream is sold in quarts (a familiar, non-SI base unit), pints
(0.5 quart), or gallons (4 quarts). We also use fractions or multiples of units in the SI system, but these fractions or multiples are
always powers of 10. Fractional or multiple SI units are named using a prefix and the name of the base unit. For example, a length
of 1000 meters is also called a kilometer because the prefix kilo means “one thousand,” which in scientific notation is 103 (1
kilometer = 1000 m = 103 m). The prefixes used and the powers to which 10 are raised are listed in Table 1.4.2.
Table 1.4.2: Common Unit Prefixes
Prefix Symbol Factor Example
SI Base Units
The initial units of the metric system, which eventually evolved into the SI system, were established in France during the French
Revolution. The original standards for the meter and the kilogram were adopted there in 1799 and eventually by other countries.
This section introduces four of the SI base units commonly used in chemistry. Other SI units will be introduced in subsequent
chapters.
Length
The standard unit of length in both the SI and original metric systems is the meter (m). A meter was originally specified as
1/10,000,000 of the distance from the North Pole to the equator. It is now defined as the distance light in a vacuum travels in
1/299,792,458 of a second. A meter is about 3 inches longer than a yard (Figure 1.4.1); one meter is about 39.37 inches or 1.094
yards. Longer distances are often reported in kilometers (1 km = 1000 m = 103 m), whereas shorter distances can be reported in
centimeters (1 cm = 0.01 m = 10−2 m) or millimeters (1 mm = 0.001 m = 10−3 m).
1.4.2 https://chem.libretexts.org/@go/page/217238
Figure 1.4.1 : The relative lengths of 1 m, 1 yd, 1 cm, and 1 in. are shown (not actual size), as well as comparisons of 2.54 cm and
1 in., and of 1 m and 1.094 yd.
Mass
The standard unit of mass in the SI system is the kilogram (kg). A kilogram was originally defined as the mass of a liter of water (a
cube of water with an edge length of exactly 0.1 meter). In 1889, it was redefined by a certain cylinder of platinum-iridium alloy,
which was kept in France (Figure 1.4.2). Any object with the same mass as this cylinder was said to have a mass of 1 kilogram
(which can lead to uncertainties unacceptable to the precision of modern instrumentation). One kilogram is about 2.2 pounds. The
gram (g) is exactly equal to 1/1000 of the mass of the kilogram (10−3 kg). Over the past 100 years, the IPK has lost 50 millionths of
a gram - a seemingly negligible amount, but something that has caused it to be lighter - or all standard replicas to be heavier - and
changing the definition of a kilogram in the process. As all balances in the world are standardized to this value, it is important that
this value, itself, be standard. On May 20, 2019, a new definition will be used for the kilogram, based on the unchanging Planck's
constant.1
Figure 1.4.2 : This replica prototype kilogram is housed at the National Institute of Standards and Technology (NIST) in
Maryland. (credit: National Institutes of Standards and Technology).
1.4.3 https://chem.libretexts.org/@go/page/217238
A Kilogram Is Now a Kilogram—Forever …
Video 1.4.2 : For more information on the new definition of the kilogram, check out this video!
Time
The SI base unit of time is the second (s). Small and large time intervals can be expressed with the appropriate prefixes; for
example, 3 microseconds = 0.000003 s = 3 × 10−6 and 5 megaseconds = 5,000,000 s = 5 × 106 s. Alternatively, hours, days, and
years can be used.
Derived SI Units
We can derive many units from the seven SI base units. For example, we can use the base unit of length to define a unit of volume,
and the base units of mass and length to define a unit of density.
Volume
Volume is the measure of the amount of space occupied by an object. The standard SI unit of volume is defined by the base unit of
length (Figure 1.4.3). The standard volume is a cubic meter (m3), a cube with an edge length of exactly one meter. To dispense a
cubic meter of water, we could build a cubic box with edge lengths of exactly one meter. This box would hold a cubic meter of
water or any other substance.
A more commonly used unit of volume is derived from the decimeter (0.1 m, or 10 cm). A cube with edge lengths of exactly one
decimeter contains a volume of one cubic decimeter (dm3). A liter (L) is the more common name for the cubic decimeter. One liter
is about 1.06 quarts. A cubic centimeter (cm3) is the volume of a cube with an edge length of exactly one centimeter. The
abbreviation cc (for cubic centimeter) is often used by health professionals. A cubic centimeter is also called a milliliter (mL) and is
1/1000 of a liter.
<
Figure 1.4.3 : (a) The relative volumes are shown for cubes of 1 m3, 1 dm3 (1 L), and 1 cm3 (1 mL) (not to scale). (b) The diameter
of a dime is compared relative to the edge length of a 1-cm3 (1-mL) cube.
1.4.4 https://chem.libretexts.org/@go/page/217238
Density
We use the mass and volume of a substance to determine its density. Thus, the units of density are defined by the base units of mass
and length.
The density of a substance is the ratio of the mass of a sample of the substance to its volume. The SI unit for density is the kilogram
per cubic meter (kg/m3). For many situations, however, this as an inconvenient unit, and we often use grams per cubic centimeter
(g/cm3) for the densities of solids and liquids, and grams per liter (g/L) for gases. Although there are exceptions, most liquids and
solids have densities that range from about 0.7 g/cm3 (the density of gasoline) to 19 g/cm3 (the density of gold). The density of air
is about 1.2 g/L. Table 1.4.3 shows the densities of some common substances.
Table 1.4.3: Densities of Common Substances
Solids Liquids Gases (at 25 °C and 1 atm)
ice (at 0 °C) 0.92 g/cm3 water 1.0 g/cm3 dry air 1.20 g/L
oak (wood) 0.60–0.90 g/cm3 ethanol 0.79 g/cm3 oxygen 1.31 g/L
copper 9.0 g/cm3 glycerin 1.26 g/cm3 carbon dioxide 1.80 g/L
lead 11.3 g/cm3 olive oil 0.92 g/cm3 helium 0.16 g/L
While there are many ways to determine the density of an object, perhaps the most straightforward method involves separately
finding the mass and volume of the object, and then dividing the mass of the sample by its volume. In the following example, the
mass is found directly by weighing, but the volume is found indirectly through length measurements.
mass
density = (1.4.1)
volume
Example 1.4.1
Calculation of Density Gold—in bricks, bars, and coins—has been a form of currency for centuries. In order to swindle people
into paying for a brick of gold without actually investing in a brick of gold, people have considered filling the centers of hollow
gold bricks with lead to fool buyers into thinking that the entire brick is gold. It does not work: Lead is a dense substance, but its
density is not as great as that of gold, 19.3 g/cm3. What is the density of lead if a cube of lead has an edge length of 2.00 cm and
a mass of 90.7 g?
Solution
The density of a substance can be calculated by dividing its mass by its volume. The volume of a cube is calculated by cubing
the edge length.
3
volume of lead cube = 2.00 cm × 2.00 cm × 2.00 cm = 8.00 cm
(We will discuss the reason for rounding to the first decimal place in the next section.)
Exercise 1.4.1
a. To three decimal places, what is the volume of a cube (cm3) with an edge length of 0.843 cm?
b. If the cube in part (a) is copper and has a mass of 5.34 g, what is the density of copper to two decimal places?
Answer a
0.599 cm3;
Answer b
8.91 g/cm3
1.4.5 https://chem.libretexts.org/@go/page/217238
Example 1.4.2 : Using Displacement of Water to Determine Density
This PhET simulation illustrates another way to determine density, using displacement of water. Determine the density of the
red and yellow blocks.
Solution
When you open the density simulation and select Same Mass, you can choose from several 5.00-kg colored blocks that you can
drop into a tank containing 100.00 L water. The yellow block floats (it is less dense than water), and the water level rises to
105.00 L. While floating, the yellow block displaces 5.00 L water, an amount equal to the weight of the block. The red block
sinks (it is more dense than water, which has density = 1.00 kg/L), and the water level rises to 101.25 L.
The red block therefore displaces 1.25 L water, an amount equal to the volume of the block. The density of the red block is:
mass 5.00 kg
density = = = 4.00 kg/L (1.4.2)
volume 1.25 L
Note that since the yellow block is not completely submerged, you cannot determine its density from this information. But if
you hold the yellow block on the bottom of the tank, the water level rises to 110.00 L, which means that it now displaces 10.00
L water, and its density can be found:
mass 5.00 kg
density = = = 0.500 kg/L (1.4.3)
volume 10.00 L
Exercise 1.4.1
Remove all of the blocks from the water and add the green block to the tank of water, placing it approximately in the middle of
the tank. Determine the density of the green block.
Answer
2.00 kg/L
Summary
Measurements provide quantitative information that is critical in studying and practicing chemistry. Each measurement has an
amount, a unit for comparison, and an uncertainty. Measurements can be represented in either decimal or scientific notation.
Scientists primarily use the SI (International System) or metric systems. We use base SI units such as meters, seconds, and
kilograms, as well as derived units, such as liters (for volume) and g/cm3 (for density). In many cases, we find it convenient to use
unit prefixes that yield fractional and multiple units, such as microseconds (10−6 seconds) and megahertz (106 hertz), respectively.
Footnotes
1. Read more about the redefinition of SI units including the kilogram here (Laura Howe, CE&N, Nov. 16, 2018).
Key Equations
mass
density =
volume
Glossary
density
ratio of mass to volume for a substance or object
kilogram (kg)
1.4.6 https://chem.libretexts.org/@go/page/217238
standard SI unit of mass; 1 kg = approximately 2.2 pounds
length
measure of one dimension of an object
liter (L)
(also, cubic decimeter) unit of volume; 1 L = 1,000 cm3
meter (m)
standard metric and SI unit of length; 1 m = approximately 1.094 yards
milliliter (mL)
1/1,000 of a liter; equal to 1 cm3
second (s)
SI unit of time
unit
standard of comparison for measurements
volume
amount of space occupied by an object
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
Sci Show is a division of Complexly and videos are free to stream for educational purposes.
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
1.4: Measurements is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
1.4.7 https://chem.libretexts.org/@go/page/217238
1.4.1: Practice Problems on Measurements
PROBLEM 1.4.1.1
Indicate the SI base units or derived units that are appropriate for the following measurements.
(a) the mass of the moon
(b) the distance from Dallas to Oklahoma City
(c) the speed of sound
(d) the density of air
(e) the area of the state of Oregon
(f) the volume of a flu shot
Answer a
kilogram
Answer b
meter
Answer c
meter per second
Answer d
kilogram per cubic meter
Answer e
square meters
Answer f
cubic meters
PROBLEM 1.4.1.2
Indicate the SI base units or derived units that are appropriate for the following measurements.
(a) the length of a marathon race (26.2 miles)
(b) the mass of an automobile
(c) the volume of a swimming pool
(d) the speed of an airplane
(e) the density of gold
(f) the area of a football field
Answer a
meter
Answer b
kilogram
Answer c
cubic meter
1.4.1.1 https://chem.libretexts.org/@go/page/217242
Answer d
meter per second
Answer e
kilogram per cubic meter
Answer f
square meter
PROBLEM 1.4.1.3
A large piece of jewelry has a mass of 132.6 g. A graduated cylinder initially contains 48.6 mL of water. When the jewelry is
submerged in the graduated cylinder, the total volume increases to 61.2 mL.
(a) Determine the density of this piece of jewelry (in g/mL).
(b) Assuming the jewelry is made of only one substance, what substance is it likely to be (Refer to Table 1.1.3 in the
previous section). Explain your answer.
Answer a
10.5 g/mL
Answer b
Silver (Density is 10.5 g/mL in Table 1.1.3)
Problem 1.1.3
PROBLEM 1.4.1.4
Give the name and symbol of the prefixes used with SI units to indicate multiplication by the following exact quantities.
(a) 103
(b) 10-2
(c) 0.1
(d) 10-3
(e) 1,000,000
(f) 0.000001
Answer a
kilo- (k)
1.4.1.2 https://chem.libretexts.org/@go/page/217242
Answer b
centi- (c)
Answer c
deci- (d)
Answer d
milli- (m)
Answer e
Mega (M)
Answer f
micro (u or µ)
PROBLEM 1.4.1.5
Give the name of the prefix and the quantity indicated by the following symbols that are used with SI base units.
(a) c
(b) d
(c) G
(d) k
(e) m
(f) n
(g) p
(h) T
Answer a
centi (10-2)
Answer b
deci (10-1)
Answer c
giga (109)
Answer d
kilo (103)
Answer e
milli (10-3)
Answer f
nano (10-9)
Answer g
pico (10-12)
Answer h
1.4.1.3 https://chem.libretexts.org/@go/page/217242
Tera (1012)
1.4.1: Practice Problems on Measurements is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
1.4.1.4 https://chem.libretexts.org/@go/page/217242
1.5: The Mole is a Measure of Amount
Skills to Develop
Define the amount unit mole and the related quantity Avogadro’s number
The Mole
The identity of a substance is defined not only by the types of atoms or ions it contains, but by the quantity of each type of atom or
ion. For example, water, H2O, and hydrogen peroxide, H2O2, are alike in that their respective molecules are composed of hydrogen
and oxygen atoms. However, because a hydrogen peroxide molecule contains two oxygen atoms, as opposed to the water molecule,
which has only one, the two substances exhibit very different properties. Today, we possess sophisticated instruments that allow the
direct measurement of these defining microscopic traits; however, the same traits were originally derived from the measurement of
macroscopic properties (the masses and volumes of bulk quantities of matter) using relatively simple tools (balances and
volumetric glassware). This experimental approach required the introduction of a new unit for amount of substances, the mole,
which remains indispensable in modern chemical science.
The mole is an amount unit similar to familiar units like pair, dozen, gross, etc. It provides a specific measure of the number of
atoms or molecules in a bulk sample of matter. A mole is defined as the amount of substance containing the same number of
discrete entities (such as atoms, molecules, and ions) as the number of atoms in a sample of pure 12C weighing exactly 12 g. One
Latin connotation for the word “mole” is “large mass” or “bulk,” which is consistent with its use as the name for this unit. The
mole provides a link between an easily measured macroscopic property, bulk mass, and an extremely important fundamental
property, number of atoms, molecules, and so forth.
The number of entities composing a mole has been experimentally determined to be 6.02214179 × 10 , a fundamental constant
23
named Avogadro’s number (NA) or the Avogadro constant in honor of Italian scientist Amedeo Avogadro. This constant is properly
reported with an explicit unit of “per mole,” a conveniently rounded version being 6.022 × 10 /mol. 23
Consistent with its definition as an amount unit, 1 mole of any element contains the same number of atoms as 1 mole of any other
element. The masses of 1 mole of different elements, however, are different, since the masses of the individual atoms are drastically
different. The molar mass of an element (or compound) is the mass in grams of 1 mole of that substance, a property expressed in
units of grams per mole (g/mol).
1.5: The Mole is a Measure of Amount is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
1.5.1 https://chem.libretexts.org/@go/page/217239
1.6: Accuracy and Precision
Figure 1.6.1 (Credit: Daniel Arizpe; Source: Commons Wikimedia, High School Basketball Game(opens in new window)
[commons.wikimedia.org]; License: Public Domain)
Figure 1.6.2 : The distribution of darts on a dartboard shows the difference between accuracy and precision. (Credit: Christopher
Auyeung; Source: CK-12 Foundation; License: CC BY-NC 3.0(opens in new window))
Assume that three darts are thrown at the dartboard, with the bulls-eye representing the true, or accepted, value of what is being measured. A dart
that hits the bulls-eye is highly accurate, whereas a dart that lands far away from the bulls-eye displays poor accuracy. The figure above
demonstrates four possible outcomes:
a. The darts have landed far from each other and far from the bulls-eye. This grouping demonstrates measurements that are neither accurate nor
precise.
b. The darts are close to one another, but far from the bulls-eye. This grouping demonstrates measurements that are precise, but not accurate. In a
laboratory situation, high precision with low accuracy often results from a systematic error. Either the measurer makes the same mistake
repeatedly, or the measuring tool is somehow flawed. A poorly calibrated balance may give the same mass reading every time, but it will be far
from the true mass of the object.
c. The darts are not grouped very near to each other, but are generally centered around the bulls-eye. This demonstrates poor precision, but fairly
high accuracy. This situation is not desirable in a lab situation because the "high" accuracy may simply be random chance and not a true indicator
of good measuring skill.
d. The darts are grouped together and have hit the bulls-eye. This demonstrates high precision and high accuracy. Scientists always strive to
maximize both in their measurements.
1.6.1 https://chem.libretexts.org/@go/page/217240
Figure 1.6.3 : Students in a chemistry lab are making careful measurements with a series of volumetric flasks. Accuracy and precision are critical in
every experiment. (Credit: CK-12 Foundation; Source: CK-12 Foundation; License: CK-12 Curriculum Materials license)
0:00
0:00
1.6.2 https://chem.libretexts.org/@go/page/217240
Summary
Accuracy is a measure of how close a measurement is to the correct or accepted value of the quantity being measured.
Precision is a measure of how close a series of measurements are to one another.
Review
1. Define accuracy.
2. Define precision.
3. What can be said about the reproducibility of precise values?
1.6: Accuracy and Precision is shared under a CC BY-NC license and was authored, remixed, and/or curated by LibreTexts.
3.12: Accuracy and Precision by CK-12 Foundation is licensed CK-12. Original source: https://flexbooks.ck12.org/cbook/ck-12-chemistry-flexbook-2.0/.
1.6.3 https://chem.libretexts.org/@go/page/217240
1.7: Dimensional Analysis
Skills to Develop
Explain the dimensional analysis (factor label) approach to mathematical calculations involving quantities
Use dimensional analysis to carry out unit conversions for a given property and computations involving two or more
properties
Perform dimensional analysis calculations with units raised to a power
Use density as a conversion factor
Note that this simple arithmetic involves dividing the numbers of each measured quantity to yield the number of the computed
quantity (100/10 = 10) and likewise dividing the units of each measured quantity to yield the unit of the computed quantity (m/s =
m/s). Now, consider using this same relation to predict the time required for a person running at this speed to travel a distance of 25
m. The same relation between the three properties is used, but in this case, the two quantities provided are a speed (10 m/s) and a
distance (25 m). To yield the sought property, time, the equation must be rearranged appropriately:
distance
time = (1.7.3)
speed
Again, arithmetic on the numbers (25/10 = 2.5) was accompanied by the same arithmetic on the units (m/m/s = s) to yield the
number and unit of the result, 2.5 s. Note that, just as for numbers, when a unit is divided by an identical unit (in this case, m/m),
the result is “1”—or, as commonly phrased, the units “cancel.”
1.7.1 https://chem.libretexts.org/@go/page/217241
These calculations are examples of a versatile mathematical approach known as dimensional analysis (or the factor-label method).
Dimensional analysis is based on this premise: the units of quantities must be subjected to the same mathematical operations as
their associated numbers. This method can be applied to computations ranging from simple unit conversions to more complex,
multi-step calculations involving several different quantities.
Several other commonly used conversion factors are given in Table 1.7.1.
Table 1.7.1: Common Conversion Factors
Length Volume Mass
1 cm3 = 1 mL
When we multiply a quantity (such as distance given in inches) by an appropriate unit conversion factor, we convert the quantity to
an equivalent value with different units (such as distance in centimeters). For example, a basketball player’s vertical jump of 34
inches can be converted to centimeters by:
2.54 cm
34 in. × = 86 cm (1.7.6)
1 in.
Since this simple arithmetic involves quantities, the premise of dimensional analysis requires that we multiply both numbers and
units. The numbers of these two quantities are multiplied to yield the number of the product quantity, 86, whereas the units are
multiplied to yield
in. ×cm
. (1.7.7)
in.
Just as for numbers, a ratio of identical units is also numerically equal to one,
in.
=1 (1.7.8)
in.
and the unit product thus simplifies to cm. (When identical units divide to yield a factor of 1, they are said to “cancel.”) Using
dimensional analysis, we can determine that a unit conversion factor has been set up correctly by checking to confirm that the
original unit will cancel, and the result will contain the sought (converted) unit.
1.7.2 https://chem.libretexts.org/@go/page/217241
If we have the conversion factor, we can determine the mass in kilograms using an equation similar the one used for converting
length from inches to centimeters.
The correct unit conversion factor is the ratio that cancels the units of grams and leaves ounces.
1 oz
x oz = 125 g ×
28.349 g
125
=( ) oz
28.349
Exercise 1.7.1
Convert a volume of 9.345 qt to liters.
Answer
8.844 L
Beyond simple unit conversions, the factor-label method can be used to solve more complex problems involving computations.
Regardless of the details, the basic approach is the same—all the factors involved in the calculation must be appropriately oriented
to insure that their labels (units) will appropriately cancel and/or combine to yield the desired unit in the result. This is why it is
referred to as the factor-label method. As your study of chemistry continues, you will encounter many opportunities to apply this
approach.
Then,
3
4.20 × 10 g
density = = 1.11 g/mL
3
3.78 × 10 mL
1.7.3 https://chem.libretexts.org/@go/page/217241
Alternatively, the calculation could be set up in a way that uses three unit conversion factors sequentially as follows:
Exercise 1.7.2
What is the volume in liters of 1.000 oz, given that 1 L = 1.0567 qt and 1 qt = 32 oz (exactly)?
Answer
−2
2.956 × 10 L
1.0567 qt 1 gal
213 L × × = 56.3 gal
1 L 4 qt
Then,
777 mi
(average) mileage = = 13.8 miles/gallon = 13.8 mpg
56.3 gal
Alternatively, the calculation could be set up in a way that uses all the conversion factors sequentially, as follows:
1250 km 0.62137 mi 1 L 4 qt
× × × = 13.8 mpg
1.0567 qt 1 gal
213 L 1 km
Exercise 1.7.3
A Toyota Prius Hybrid uses 59.7 L gasoline to drive from San Francisco to Seattle, a distance of 1300 km (two significant
digits).
a. What (average) fuel economy, in miles per gallon, did the Prius get during this trip?
b. If gasoline costs $3.90 per gallon, what was the fuel cost for this trip?
Answer a
51 mpg
Answer b
$62
1.7.4 https://chem.libretexts.org/@go/page/217241
Units Raised to a Power
When calculating area or volume, you multiply together lengths, widths, and heights. Just like in our dimensional analysis above,
our units and our numbers both undergo the mathematical operation, meaning that multiplying the quantity of length by the
quantity of width also multiplies the units.
2
This complicates the conversion of units, however, since our GIVEN conversion factors often only account for one dimension, not
two or three. When building a conversion factor for units raised to a power, we simply raise the conversion factor to the power we
want our final units in. The trick is to remember to raise the QUANTITY and UNITS by the power.
Now, if we examine the table of conversion factors (Table 1.7.1), we find that there is 16.4 cm3 in 1 in3.
3
1in
3 3
5700c m ∗ = 347.6c m (1.7.11)
3
16.4cm
We could have just as easily have done this if we hadn't been given the direct conversion factor between cm3 and in3. We simply
would have had to raise the conversion factor between cm and in to the third power.
3
3
1in 3
5700c m ∗( ) = 347.6i n (1.7.12)
2.54cm
The trick with this way of doing the calculation is you have to remember to apply the power to EVERYTHING:
3 3 3
1in (1 i n )
( ) = (1.7.13)
3
2.54cm 2.54 c m3
Note: We are ignoring a concept known as "significant figures" in this example. You will cover the rules for significant figures
in next week's lab.
Exercise 1.7.4
Convert 274 m2 to in2.
Answer
424701 in2
Exercise 1.7.5
A commercial jet is fueled with 156,874 L of jet fuel.
a) If the density of the fuel is 0.768 g/cm3, what is the mass of the fuel in kilograms?
b) If the jet weights 443.613 Mg without passengers or fuel, what is the mass when the fuel is added?
Answer a
1.7.5 https://chem.libretexts.org/@go/page/217241
120,479 kg
Answer b
564,092 kg or 564.092 Mg
Summary
Measurements are made using a variety of units. It is often useful or necessary to convert a measured quantity from one unit into
another. These conversions are accomplished using unit conversion factors, which are derived by simple applications of a
mathematical approach called the factor-label method or dimensional analysis. This strategy is also employed to calculate sought
quantities using measured quantities and appropriate mathematical relations.
Glossary
dimensional analysis
(also, factor-label method) versatile mathematical approach that can be applied to computations ranging from simple unit
conversions to more complex, multi-step calculations involving several different quantities
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Crash Course Chemistry, Crash Course is a division of Complexly and videos are free to stream for educational purposes.
1.7: Dimensional Analysis is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
1.7.6 https://chem.libretexts.org/@go/page/217241
1.7.1: Practice Problems on Dimensional Analysis
PROBLEM 1.7.1.1
Write the conversion factors (as ratios) for the number of:
(a) kilometers in 1 mile
(b) liters in 1 liquid quart
(c) grams in 1 ounce
Answer a
1.6093 km : 1 mi
Answer b
0.94635 L : 1 qt
Answer c
28.35 g : 1 oz
PROBLEM 1.7.1.2
The label on a soft drink bottle gives the volume in two units: 2.0 L and 67.6 fl oz. Use this information to derive a conversion
factor between the English and metric units.
Answer
2.0L 0.030L
= (1.7.1.1)
67.6f loz. 1f loz.
PROBLEM 1.7.1.3
The label on a box of cereal gives the mass of cereal in two units: 978 grams and 34.5 oz. Use this information to find a
conversion factor between the English and metric units.
Answer
1.7.1.1 https://chem.libretexts.org/@go/page/217243
978g 28.35g
= (1.7.1.2)
34.5oz 1oz.
PROBLEM 1.7.1.4
Soccer is played with a round ball having a circumference between 27 and 28 inches and a mass between 14 and 16 oz. What
are these specifications in cm and g?
Answer
68.6 cm to 71.1 cm
396.9 g to 453.6 g
PROBLEM 1.7.1.5
How many milliliters are in a 12 oz soda can?
Answer
354.9 mL
PROBLEM 1.7.1.6
A barrel of oil is exactly 42 gal. How many liters of oil are in the barrel?
Answer
159 L
PROBLEM 1.7.1.7
The diameter of a red blood cell is about 3 x 10-4 inches. What is the diameter in centimeters?
Answer
7.6 x 10-4 cm
PROBLEM 1.7.1.8
The distance between the centers of two oxygen atoms in an oxygen molecule is 1.21 x 10-8 cm. What is this distance in inches?
1.7.1.2 https://chem.libretexts.org/@go/page/217243
Answer
4.76 x 10-9 in
PROBLEM 1.7.1.9
Is a 197-lb weight lifter light enough to compete in a class limited to those weighing 90 kg or less?
Answer
Yes (They weigh 89.35 kg)
PROBLEM 1.7.1.10
Complete the following conversions between SI units.
(a) 612 g = ? mg
(b) 8.160 m = ? cm
(c) 3779 µg = ? g
(d) 781 mL = ? L
(e) 4.18 kg = ? g
(f) 27.8 m = ? km
(g) 0.13 mL = ? L
(h) 1738 km = ? m
(i) 1.9 Gg = ? g
Answer a
612,000 mg
Answer b
816.0 cm
Answer c
3.779 x 10-3 g
Answer d
0.781 L
Answer e
4180 g
Answer f
0.0278 km
1.7.1.3 https://chem.libretexts.org/@go/page/217243
Answer g
1.3 x 10-2 L
Answer h
1,738,000 m
Answer i
1.9 x 109 g
PROBLEM 1.7.1.11
Make the conversion indicated in each of the following:
(a) the men's world record long jump, 29 ft 4.5 in, to meters
(b) the greatest depth of the ocean, about 6.5 mi, to kilometers
(c) the area of an 8.5 by 11 inch sheet of paper in cm2
(d) The displacement volume of an automobile engine, 161 in3, to L
(e) the estimated mass of the atmosphere, 5.6 x 1015 tons, to kilograms (1 ton = 2000 lbs)
(f) the mass of a bushel of rye, 32.0 lb, to kilograms
(g) the mass of a 5.00 grain aspirin tablet to milligrams (1 grain = 0.00229 oz)
Answer a
8.96 m
Answer b
10.46 km
Answer c
603.22 cm2
Answer d
2.64 L
Answer e
5.08 x 1018 kg
Answer f
14.52 kg
Answer g
324 mg
1.7.1.4 https://chem.libretexts.org/@go/page/217243
Click here to see a video of the solution(s).
Problem 1.2.11
PROBLEM 1.7.1.12
Many chemistry conferences have held a 50-Trillion Angstrom (Å) Run. How long is this run in kilometers and in miles? (1 Å =
1 x 10-10 m)
Answer
5 kilometers or 3.1 miles
PROBLEM 1.7.1.13
As an instructor is preparing for an experiment, he requires 225 g phosphoric acid. The only container readily available is a 150-
mL Erlenmeyer flask. Is it large enough to contain the acid, the density of which is 1.83 g/mL?
Answer
Yes, because the acid's volume will be 122.95 mL
PROBLEM 1.7.1.14
In a recent Grand Prix, the winner completed the race with an average speed of 229.8 km/h. What was the speed in miles per
hour, meters per second, and feet per second?
Answer
142.8 mi/h; 63.8 m/s; 209 ft/s
1.7.1.5 https://chem.libretexts.org/@go/page/217243
PROBLEM 1.7.1.15
Calculate these masses.
(a) what is the mass of 6.00 cm3 of mercury (density = 13.5939 g/cm3)?
(b) what is the mass of 25.0 mL octane (density = 0.702 g/cm3)?
(c) what is the mass of 4.00 cm3 of sodium (density = 0.97 g/cm3)?
(d) What is the mass of 125 mL gaseous chlorine (density = 3.16 g/L)?
Answer a
81.5634 g
Answer b
17.55 g
Answer c
3.88 g
Answer d
0.395 g
Problem 1.2.15
PROBLEM 1.7.1.16
Calculate the following volumes.
1.7.1.6 https://chem.libretexts.org/@go/page/217243
(a) What is the volume of 25 g of iodine (density = 4.93 g/cm3)?
(b) What is the volume of 3.28 g gaseous hydrogen (density = 0.089 g/L)?
(c) What is the volume of 11.3 g graphite (density = 2.25 g/cm3)?
(d) What is the volume of 39.657 g bromine (density = 2.928 g/cm3)?
Answer a
5.07 mL
Answer b
36.9 L
Answer c
5.02 mL
Answer d
13.54 mL
PROBLEM 1.7.1.17
Convert 195.7 in2 to m2.
Answer
0.126 m2
1.7.1: Practice Problems on Dimensional Analysis is shared under a not declared license and was authored, remixed, and/or curated by
LibreTexts.
1.7.1.7 https://chem.libretexts.org/@go/page/217243
1.8: Significant Digits
Accuracy and precision are very important in chemistry. However, the laboratory equipment and machines used in labs are limited
in such a way that they can only determine a certain amount of data. For example, a scale can only mass an object up until a certain
decimal place, because no machine is advanced enough to determine an infinite amount of digits. Machines are only able to
determine a certain amount of digits precisely. These numbers that are determined precisely are called significant digits. Thus, a
scale that could only mass until 99.999 mg, could only measure up to 5 figures of accuracy (5 significant digits). Furthermore, in
order to have accurate calculations, the end calculation should not have more significant digits than the original set of data.
Introduction
Significant Digits - Number of digits in a figure that express the precision of a measurement instead of its magnitude. The easiest
method to determine significant digits is done by first determining whether or not a number has a decimal point. This rule is known
as the Atlantic-Pacific Rule. The rule states that if a decimal point is Absent, then the zeroes on the Atlantic/right side are
insignificant. If a decimal point is Present, then the zeroes on the Pacific/left side are insignificant.
Example 1.8.1 :
The first two zeroes in 200500 (four significant digits) are significant because they are between two non-zero digits, and the last
two zeroes are insignificant because they are after the last non-zero digit.
It should be noted that both constants and quantities of real world objects have an infinite number of significant figures. For
example if you were to count three oranges, a real world object, the value three would be considered to have an infinite number
of significant figures in this context.
Example 1.8.1
How many significant digits are in 5010?
Solution
1. Start counting for significant digits On the first non-zero digit (5).
2. Stop counting for significant digits On the last non-zero digit (1).
5 0 1 0 Key: 0 = significant zero. 0 = insignificant zero.
3 significant digits.
Example 1.8.3
1.8.1 https://chem.libretexts.org/@go/page/242394
The first two zeroes in 0.058000 (five significant digits) are insignificant because they are before the first non-zero digit, and the
last three zeroes are significant because they are after the first non-zero digit.
Example 1.8.4
How many significant digits are in 0.70620?
Solution
1. Start counting for significant digits On the first non-zero digit (7).
2. Stop counting for significant digits On the last digit (0).
0 . 7 0 6 2 0 Key: 0 = significant zero.0 = insignificant zero.
5 significant digits.
Scientific Notation
Scientific notation form: a x 10b, where “a” and “b” are integers, and "a" has to be between 1 and 10.
Example 1.8.5
The scientific notation for 4548 is 4.548 x 103.
Solution
Disregard the “10b,” and determine the significant digits in “a.”
4.548 x 103 has 4 significant digits.
Example 1.8.6
How many significant digits are in 1.52 x 106?
NOTE: Only determine the amount of significant digits in the "1.52" part of the scientific notation form.
Answer
3 significant digits.
Example 1.8.7
Round 32445.34 to 2 significant digits.
Answer
32000 (NOT 32000.00, which has 7 significant digits. Due to the decimal point, the zeroes after the first non-zero digit become
significant).
Example 1.8.8
1.8.2 https://chem.libretexts.org/@go/page/242394
Y = 232.234 + 0.27 Find Y.
Answer
Y = 232.50
NOTE: 232.234 has 3 decimal places and 0.27 has 2 decimal places. The least amount of decimal places is 2. Thus, the answer
must be rounded to the 2nd decimal place (thousandth).
Example 1.8.9
Y = 28 x 47.3 Find Y
Answer
Y = 1300
NOTE: 28 has 2 significant digits and 47.3 has 3 significant digits. The least amount of significant digits is 2. Thus, the answer
must me rounded to 2 significant digits (which is done by keeping 2 significant digits and replacing the rest of the digits with
insignificant zeroes).
Exact Numbers
Exact numbers can be considered to have an unlimited number of significant figures, as such calculations are not subject to errors
in measurement. This may occur:
1. By definition (1 minute = 60 seconds, 1 inch = 2.54 cm, 12 inches = 1 foot, etc.)
2. As a result of counting (6 faces on a cube or dice, two hydrogen atoms in a water molecule, 3 peas in a pod, etc.)
References
1. Brown, Theodore E., H. Eugene LeMay, and Bruce E. Bursten. Chemistry: The Central Science, Tenth Edition. Pearson
Education Inc. Upper Saddle River, New Jersey: 2005.
2. Petrucci, Ralph H., William S. Harwood, F. Geoffrey Herring, and Jeffry D. Madura. General Chemistry: Principles and
Modern Applications, Ninth Edition. Pearson Education Inc. Upper Saddle River, New Jersey: 2007.
3. Petrucci, Ralph H., William S. Harwood, F. Geoffrey Herring, and Jeffry D. Madura. General Chemistry: Principles and
Modern Applications, Tenth Edition. Pearson Education Inc. Upper Saddle River, New Jersey: 2011. Custom Edition for Chem
2, University of California, Davis
Additional Problems
1. a. a) How many significant digits Are in 50?
b. b) How many significant digits Are in 50.0?
2. How many significant digits Are in 3.670 × 10 ?35
1.8.3 https://chem.libretexts.org/@go/page/242394
Solutions
1. a) 1 significant digit.
b) 2 significant digits.
2. 4 significant digits.
3. 4280000
4. 0.06
5. Y = 61.9
6. Y = -3
7. Y = 9270
8. Y = 16
9. Y = (23.2 + 16.723) x 28
Y = 1100 (NOTE: 28 has the least amount of significant digits (2 sig. figs.) Thus, answer must be rounded to 2 sig.
figs.)
Contributors
Jeffrey Susila (UCD), Neema Shah (UCD)
1.8: Significant Digits is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
1.8.4 https://chem.libretexts.org/@go/page/242394
CHAPTER OVERVIEW
2: Atoms and Elements
Unit 2 Objectives
By the end of this unit, you will be able to:
State the postulates of Dalton’s atomic theory
Use postulates of Dalton’s atomic theory to explain the laws of definite and multiple proportions
Outline milestones in the development of modern atomic theory
Summarize and interpret the results of the experiments of Thomson, Millikan, and Rutherford
Describe the three subatomic particles that compose atoms
Define isotopes and give examples for several elements
Write and interpret symbols that depict the atomic number, mass number, and charge of an atom or ion
Define the atomic mass unit and average atomic mass
Calculate average atomic mass and isotopic abundance
2: Atoms and Elements is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
1
2.1: A History of Atomic Theory
Skills to Develop
By the end of this section, you will be able to:
State the postulates of Dalton’s atomic theory
Use postulates of Dalton’s atomic theory to explain the laws of definite and multiple proportions
Outline milestones in the development of modern atomic theory
Summarize and interpret the results of the experiments of Thomson, Millikan, and Rutherford
A Video Introduction to Atomic Theory through the Nineteenth Century From Crash Course Chemistry
Video 2.1.1 : Lavoisier's discovery of The Law of Conservation of Matter led to the Laws of Definite and Multiple
Proportions and eventually Dalton's Atomic Theory.
2.1.1 https://chem.libretexts.org/@go/page/217245
Figure 2.1.1 : A pre-1982 copper penny (left) contains approximately 3 × 1022 copper atoms (several dozen are represented as
brown spheres at the right), each of which has the same chemical properties. (credit: modification of work by “slgckgc”/Flickr)
Figure 2.1.2 : Copper(II) oxide, a powdery, black compound, results from the combination of two types of atoms—copper (brown
spheres) and oxygen (red spheres)—in a 1:1 ratio. (credit: modification of work by “Chemicalinterest”/Wikimedia Commons)
Figure 2.1.3 : When the elements copper (a shiny, red-brown solid, shown here as brown spheres) and oxygen (a clear and
colorless gas, shown here as red spheres) react, their atoms rearrange to form a compound containing copper and oxygen (a
powdery, black solid). (credit copper: modification of work by http://images-of-elements.com/copper.php).
Dalton’s atomic theory provides a microscopic explanation of the many macroscopic properties of matter that you’ve learned about.
For example, if an element such as copper consists of only one kind of atom, then it cannot be broken down into simpler
substances, that is, into substances composed of fewer types of atoms. And if atoms are neither created nor destroyed during a
chemical change, then the total mass of matter present when matter changes from one type to another will remain constant (the law
of conservation of matter (or mass)).
2.1.2 https://chem.libretexts.org/@go/page/217245
The law of conservation of mass - Todd …
Solution
The starting materials consist of two green spheres and two purple spheres. The products consist of only one green sphere and
one purple sphere. This violates Dalton’s postulate that atoms are neither created nor destroyed during a chemical change, but
are merely redistributed. (In this case, atoms appear to have been destroyed.)
Exercise 2.1.1
In the following drawing, the green spheres represent atoms of a certain element. The purple spheres represent atoms of another
element. If the spheres touch, they are part of a single unit of a compound. Does the following chemical change represented by
these symbols violate any of the ideas of Dalton’s atomic theory? If so, which one
Answer
The starting materials consist of four green spheres and two purple spheres. The products consist of four green spheres and
two purple spheres. This does not violate any of Dalton’s postulates: Atoms are neither created nor destroyed, but are
redistributed in small, whole-number ratios.
Dalton knew of the experiments of French chemist Joseph Proust, who demonstrated that all samples of a pure compound contain
the same elements in the same proportion by mass. This statement is known as the law of definite proportions or the law of constant
composition. The suggestion that the numbers of atoms of the elements in a given compound always exist in the same ratio is
consistent with these observations. For example, when different samples of isooctane (a component of gasoline and one of the
standards used in the octane rating system) are analyzed, they are found to have a carbon-to-hydrogen mass ratio of 5.33:1, as
shown in Table 2.1.1.
Table 2.1.1: Constant Composition of Isooctane
2.1.3 https://chem.libretexts.org/@go/page/217245
Sample Carbon Hydrogen Mass Ratio
It is worth noting that although all samples of a particular compound have the same mass ratio, the converse is not true in general.
That is, samples that have the same mass ratio are not necessarily the same substance. For example, there are many compounds
other than isooctane that also have a carbon-to-hydrogen mass ratio of 5.33:1.00.
Dalton also used data from Proust, as well as results from his own experiments, to formulate another interesting law. The law of
multiple proportions states that when two elements react to form more than one compound, a fixed mass of one element will react
with masses of the other element in a ratio of small, whole numbers. For example, copper and chlorine can form a green, crystalline
solid with a mass ratio of 0.558 g chlorine to 1 g copper, as well as a brown crystalline solid with a mass ratio of 1.116 g chlorine to
1 g copper. These ratios by themselves may not seem particularly interesting or informative; however, if we take a ratio of these
ratios, we obtain a useful and possibly surprising result: a small, whole-number ratio.
1.116 g Cl
1 g Cu 2
= (2.1.1)
0.558 g Cl 1
1 g Cu
This 2-to-1 ratio means that the brown compound has twice the amount of chlorine per amount of copper as the green compound.
This can be explained by atomic theory if the copper-to-chlorine ratio in the brown compound is 1 copper atom to 2 chlorine atoms,
and the ratio in the green compound is 1 copper atom to 1 chlorine atom. The ratio of chlorine atoms (and thus the ratio of their
masses) is therefore 2 to 1 (Figure 2.1.4).
Figure 2.1.4 : Compared to the copper chlorine compound in (a), where copper is represented by brown spheres and chlorine by
green spheres, the copper chlorine compound in (b) has twice as many chlorine atoms per copper atom. (credit a: modification of
work by “Benjah-bmm27”/Wikimedia Commons; credit b: modification of work by “Walkerma”/Wikimedia Commons)
1 g C
2.1.4 https://chem.libretexts.org/@go/page/217245
In compound B, the mass ratio of carbon to oxygen is:
2.67 g O
1 g C
1 g C 1
=
2.67 g O 2
1 g C
This supports the law of multiple proportions. This means that A and B are different compounds, with A having one-half as
much carbon per amount of oxygen (or twice as much oxygen per amount of carbon) as B. A possible pair of compounds that
would fit this relationship would be A = CO2 and B = CO.
Exercise 2.1.2
A sample of compound X (a clear, colorless, combustible liquid with a noticeable odor) is analyzed and found to contain 14.13 g
carbon and 2.96 g hydrogen. A sample of compound Y (a clear, colorless, combustible liquid with a noticeable odor that is
slightly different from X’s odor) is analyzed and found to contain 19.91 g carbon and 3.34 g hydrogen. Are these data an
example of the law of definite proportions, the law of multiple proportions, or neither? What do these data tell you about
substances X and Y?
Answer
14.13 g C
In compound X, the mass ratio of carbon to hydrogen is .
2.96 g H
19.91 g C
In compound Y, the mass ratio of carbon to oxygen is .
3.34 g H
3.34 g H
This small, whole-number ratio supports the law of multiple proportions. This means that X and Y are different compounds.
In the two centuries since Dalton developed his ideas, scientists have made significant progress in furthering our understanding of
atomic theory. Much of this came from the results of several seminal experiments that revealed the details of the internal structure
of atoms. Here, we will discuss some of those key developments, with an emphasis on application of the scientific method, as well
as understanding how the experimental evidence was analyzed. While the historical persons and dates behind these experiments
can be quite interesting, it is most important to understand the concepts resulting from their work.
2.1.5 https://chem.libretexts.org/@go/page/217245
Figure 2.1.5 : (a) J. J. Thomson produced a visible beam in a cathode ray tube. (b) This is an early cathode ray tube, invented in
1897 by Ferdinand Braun. (c) In the cathode ray, the beam (shown in yellow) comes from the cathode and is accelerated past the
anode toward a fluorescent scale at the end of the tube. Simultaneous deflections by applied electric and magnetic fields permitted
Thomson to calculate the mass-to-charge ratio of the particles composing the cathode ray. (credit a: modification of work by Nobel
Foundation; credit b: modification of work by Eugen Nesper; credit c: modification of work by “Kurzon”/Wikimedia Commons).
Based on his observations, here is what Thomson proposed and why: The particles are attracted by positive (+) charges and
repelled by negative (−) charges, so they must be negatively charged (like charges repel and unlike charges attract); they are less
massive than atoms and indistinguishable, regardless of the source material, so they must be fundamental, subatomic constituents
of all atoms. Although controversial at the time, Thomson’s idea was gradually accepted, and his cathode ray particle is what we
now call an electron, a negatively charged, subatomic particle with a mass more than one thousand-times less that of an atom. The
term “electron” was coined in 1891 by Irish physicist George Stoney, from “electric ion.”
In 1909, more information about the electron was uncovered by American physicist Robert A. Millikan via his “oil drop”
experiments. Millikan created microscopic oil droplets, which could be electrically charged by friction as they formed or by using
X-rays. These droplets initially fell due to gravity, but their downward progress could be slowed or even reversed by an electric
field lower in the apparatus. By adjusting the electric field strength and making careful measurements and appropriate calculations,
Millikan was able to determine the charge on individual drops (Figure 2.1.2).
Figure 2.1.6 : Millikan’s experiment measured the charge of individual oil drops. The tabulated data are examples of a few
possible values.
Looking at the charge data that Millikan gathered, you may have recognized that the charge of an oil droplet is always a multiple of
a specific charge, 1.6 × 10−19 C. Millikan concluded that this value must therefore be a fundamental charge—the charge of a single
2.1.6 https://chem.libretexts.org/@go/page/217245
electron—with his measured charges due to an excess of one electron (1 times 1.6 × 10−19 C), two electrons (2 times 1.6 × 10−19
C), three electrons (3 times 1.6 × 10−19 C), and so on, on a given oil droplet. Since the charge of an electron was now known due
to Millikan’s research, and the charge-to-mass ratio was already known due to Thomson’s research (1.759 × 1011 C/kg), it only
required a simple calculation to determine the mass of the electron as well.
1 kg
−19 −31
Mass of electron = 1.602 × 10 C × = 9.107 × 10 kg (2.3.1)
11
1.759 × 10 C
Scientists had now established that the atom was not indivisible as Dalton had believed, and due to the work of Thomson, Millikan,
and others, the charge and mass of the negative, subatomic particles—the electrons—were known. However, the positively charged
part of an atom was not yet well understood. In 1904, Thomson proposed the “plum pudding” model of atoms, which described a
positively charged mass with an equal amount of negative charge in the form of electrons embedded in it, since all atoms are
electrically neutral. A competing model had been proposed in 1903 by Hantaro Nagaoka, who postulated a Saturn-like atom,
consisting of a positively charged sphere surrounded by a halo of electrons (Figure 2.1.3).
Figure 2.1.7 : (a) Thomson suggested that atoms resembled plum pudding, an English dessert consisting of moist cake with
embedded raisins (“plums”). (b) Nagaoka proposed that atoms resembled the planet Saturn, with a ring of electrons surrounding a
positive “planet.” (credit a: modification of work by “Man vyi”/Wikimedia Commons; credit b: modification of work by
“NASA”/Wikimedia Commons).
The next major development in understanding the atom came from Ernest Rutherford, a physicist from New Zealand who largely
spent his scientific career in Canada and England. He performed a series of experiments using a beam of high-speed, positively
charged alpha particles (α particles) that were produced by the radioactive decay of radium; α particles consist of two protons and
two neutrons (you will learn more about radioactive decay in the chapter on nuclear chemistry). Rutherford and his colleagues
Hans Geiger (later famous for the Geiger counter) and Ernest Marsden aimed a beam of α particles, the source of which was
embedded in a lead block to absorb most of the radiation, at a very thin piece of gold foil and examined the resultant scattering of
the α particles using a luminescent screen that glowed briefly where hit by an α particle.
What did they discover? Most particles passed right through the foil without being deflected at all. However, some were diverted
slightly, and a very small number were deflected almost straight back toward the source (Figure 2.1.4). Rutherford described
finding these results: “It was quite the most incredible event that has ever happened to me in my life. It was almost as incredible as
if you fired a 15-inch shell at a piece of tissue paper and it came back and hit you”1 (p. 68).
Figure 2.1.8 : Geiger and Rutherford fired α particles at a piece of gold foil and detected where those particles went, as shown in
this schematic diagram of their experiment. Most of the particles passed straight through the foil, but a few were deflected slightly
and a very small number were significantly deflected.
2.1.7 https://chem.libretexts.org/@go/page/217245
Here is what Rutherford deduced: Because most of the fast-moving α particles passed through the gold atoms undeflected, they
must have traveled through essentially empty space inside the atom. Alpha particles are positively charged, so deflections arose
when they encountered another positive charge (like charges repel each other). Since like charges repel one another, the few
positively charged α particles that changed paths abruptly must have hit, or closely approached, another body that also had a highly
concentrated, positive charge. Since the deflections occurred a small fraction of the time, this charge only occupied a small amount
of the space in the gold foil. Analyzing a series of such experiments in detail, Rutherford drew two conclusions:
1. The volume occupied by an atom must consist of a large amount of empty space.
2. A small, relatively heavy, positively charged body, the nucleus, must be at the center of each atom.
This analysis led Rutherford to propose a model in which an atom consists of a very small, positively charged nucleus, in which
most of the mass of the atom is concentrated, surrounded by the negatively charged electrons, so that the atom is electrically neutral
(Figure 2.1.5).
Figure 2.1.9 : The α particles are deflected only when they collide with or pass close to the much heavier, positively charged gold
nucleus. Because the nucleus is very small compared to the size of an atom, very few α particles are deflected. Most pass through
the relatively large region occupied by electrons, which are too light to deflect the rapidly moving particles.
After many more experiments, Rutherford also discovered that the nuclei of other elements contain the hydrogen nucleus as a
“building block,” and he named this more fundamental particle the proton, the positively charged, subatomic particle found in the
nucleus. With one addition, which you will learn next, this nuclear model of the atom, proposed over a century ago, is still used
today.
2.1.8 https://chem.libretexts.org/@go/page/217245
Rutherford Scatter
Plum Pudding
Rutherford Atom
Another important finding was the discovery of isotopes. During the early 1900s, scientists identified several substances that
appeared to be new elements, isolating them from radioactive ores. For example, a “new element” produced by the radioactive
decay of thorium was initially given the name mesothorium. However, a more detailed analysis showed that mesothorium was
chemically identical to radium (another decay product), despite having a different atomic mass. This result, along with similar
findings for other elements, led the English chemist Frederick Soddy to realize that an element could have types of atoms with
different masses that were chemically indistinguishable. These different types are called isotopes—atoms of the same element that
differ in mass. Soddy was awarded the Nobel Prize in Chemistry in 1921 for this discovery.
One puzzle remained: The nucleus was known to contain almost all of the mass of an atom, with the number of protons only
providing half, or less, of that mass. Different proposals were made to explain what constituted the remaining mass, including the
existence of neutral particles in the nucleus. As you might expect, detecting uncharged particles is very challenging, and it was not
until 1932 that James Chadwick found evidence of neutrons, uncharged, subatomic particles with a mass approximately the same as
that of protons. The existence of the neutron also explained isotopes: They differ in mass because they have different numbers of
neutrons, but they are chemically identical because they have the same number of protons. This will be explained in more detail
later in this unit.
2.1.9 https://chem.libretexts.org/@go/page/217245
The Nucleus: Crash Course Chemistry #1
Summary
2.1.10 https://chem.libretexts.org/@go/page/217245
Although no one has actually seen the inside of an atom, experiments have demonstrated much about atomic structure. Thomson’s
cathode ray tube showed that atoms contain small, negatively charged particles called electrons. Millikan discovered that there is a
fundamental electric charge—the charge of an electron. Rutherford’s gold foil experiment showed that atoms have a small, dense,
positively charged nucleus; the positively charged particles within the nucleus are called protons. Chadwick discovered that the
nucleus also contains neutral particles called neutrons. Soddy demonstrated that atoms of the same element can differ in mass;
these are called isotopes.
Footnotes
1. Ernest Rutherford, “The Development of the Theory of Atomic Structure,” ed. J. A. Ratcliffe, in Background to Modern
Science, eds. Joseph Needham and Walter Pagel, (Cambridge, UK: Cambridge University Press, 1938), 61–74. Accessed
September 22, 2014, https://ia600508.us.archive.org/3/it...e032734mbp.pdf.
Glossary
Dalton’s atomic theory
set of postulates that established the fundamental properties of atoms
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
TED-Ed’s commitment to creating lessons worth sharing is an extension of TED’s mission of spreading great ideas. Within
TED-Ed’s growing library of TED-Ed animations, you will find carefully curated educational videos, many of which
represent collaborations between talented educators and animators nominated through the TED-Ed website.
2.1: A History of Atomic Theory is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
2.1.11 https://chem.libretexts.org/@go/page/217245
2.1.1: Practice Problems- A History of Atomic Theory (Optional)
PROBLEM 2.1.1.1
In the following drawing, the green spheres represent atoms of a certain element. The purple spheres represent atoms of another
element. If the spheres of different elements touch, they are part of a single unit of a compound. The following chemical change
represented by these spheres may violate one of the ideas of Dalton's atomic theory. Which one?
Answer
The starting materials consist of one green sphere and two purple spheres. The products consist of two green spheres and
two purple spheres. This violates Dalton’s postulate that that atoms are not created during a chemical change, but are merely
redistributed.
PROBLEM 2.1.1.2
Which postulate of Dalton's theory is consistent with the following observation concerning the weights of reactants and
products?
When 100 grams of solid calcium carbonate is heated, 44 g of CO2 and 56 g of CaO are produced.
Answer
Atoms are neither created nor destroyed during a chemical change, but are instead rearranged to yield substances that are
different from those present before the change (Based on the Law of Conservation of Mass).
PROBLEM 2.1.1.3
Samples of compound X, Y, and Z are analyzed, with results shown here. Do these data provide example(s) of the law of
definite proportions, the law of multiple proportions, neither, or both? What do these data tell you about compounds X, Y, and
Z?
Answer
X+Z are similar compounds (same ratios of C and H), aligning with the Law of Definite Proportions
X+Y and Y+Z are different compounds (differing ratios of C and H), aligning with the Law of Multiple Proportions
2.1.1.1 https://chem.libretexts.org/@go/page/217249
Problem 2.1.3
PROBLEM 2.1.1.4
How are electrons and protons similar? How are they different?
Answer
Electrons and protons are both charged subatomic particles.
Protons are much larger than electrons (contributing more mass to the overall atom).
Changing the number of protons changes the identity of the atom, which changing the number of electrons changes the
charge.
PROBLEM 2.1.1.5
How are protons and neutrons similar? How are they different?
Answer
Protons and neutrons are both located in the nucleus of the atom.
Protons and neutrons both contribute to the overall mass of the atom.
Protons carry a charge while neutrons are neutral.
PROBLEM 2.1.1.6
Predict and test the behavior of α particles fired at a “plum pudding” model atom.
(a) Predict the paths taken by α particles that are fired at atoms with a Thomson’s plum pudding model structure. Explain
why you expect the α particles to take these paths.
(b) If α particles of higher energy than those in (a) are fired at plum pudding atoms, predict how their paths will differ
from the lower-energy α particle paths. Explain your reasoning.
(c) Now test your predictions from (a) and (b).
2.1.1.2 https://chem.libretexts.org/@go/page/217249
Rutherford Scatter
Plum Pudding
Rutherford Atom
Select the “Plum Pudding Atom” tab above. Set “Alpha Particles Energy” to “min,” and select “show traces.” Click on the gun
to start firing α particles. Does this match your prediction from (a)? If not, explain why the actual path would be that shown in
the simulation. Hit the pause button, or “Reset All.” Set “Alpha Particles Energy” to “max,” and start firing α particles. Does
this match your prediction from (b)? If not, explain the effect of increased energy on the actual paths as shown in the simulation.
Answer a
The plum pudding model indicates that the positive charge is spread uniformly throughout the atom, so we expect the α
particles to (perhaps) be slowed somewhat by the positive-positive repulsion, but to follow straight-line paths (i.e., not to be
deflected) as they pass through the atoms.
Answer b
Higher-energy α particles will be traveling faster (and perhaps slowed less) and will also follow straight-line paths through
the atoms.
2.1.1.3 https://chem.libretexts.org/@go/page/217249
Answer c
The α particles followed straight-line paths through the plum pudding atom. There was no apparent slowing of the α particles
as they passed through the atoms.
PROBLEM 2.1.1.7
Predict and test the behavior of α particles fired at a Rutherford atom model.
(a) Predict the paths taken by α particles that are fired at atoms with a Rutherford atom model structure. Explain why you
expect the α particles to take these paths.
(b) If α particles of higher energy than those in (a) are fired at Rutherford atoms, predict how their paths will differ from
the lower-energy α particle paths. Explain your reasoning.
(c) Predict how the paths taken by the α particles will differ if they are fired at Rutherford atoms of elements other than
gold. What factor do you expect to cause this difference in paths, and why?
(d) Now test your predictions from (a), (b), and (c).
Rutherford Scatter
Plum Pudding
Rutherford Atom
2.1.1.4 https://chem.libretexts.org/@go/page/217249
Select the “Rutherford Atom” tab above. Due to the scale of the simulation, it is best to start with a small nucleus, so
select “20” for both protons and neutrons, “min” for energy, show traces, and then start firing α particles. Does this match
your prediction from (a)? If not, explain why the actual path would be that shown in the simulation. Pause or reset, set
energy to “max,” and start firing α particles. Does this match your prediction from (b)? If not, explain the effect of
increased energy on the actual path as shown in the simulation. Pause or reset, select “40” for both protons and neutrons,
“min” for energy, show traces, and fire away. Does this match your prediction from (c)? If not, explain why the actual
path would be that shown in the simulation. Repeat this with larger numbers of protons and neutrons. What generalization
can you make regarding the type of atom and effect on the path of α particles? Be clear and specific.
Answer a
The Rutherford atom has a small, positively charged nucleus, so most α particles will pass through empty space far from the
nucleus and be undeflected. Those α particles that pass near the nucleus will be deflected from their paths due to positive-
positive repulsion. The more directly toward the nucleus the α particles are headed, the larger the deflection angle will be.
Answer b
Higher-energy α particles that pass near the nucleus will still undergo deflection, but the faster they travel, the less the
expected angle of deflection
Answer c
If the nucleus is smaller, the positive charge is smaller and the expected deflections are smaller—both in terms of how
closely the α particles pass by the nucleus undeflected and the angle of deflection. If the nucleus is larger, the positive charge
is larger and the expected deflections are larger—more α particles will be deflected, and the deflection angles will be larger.
Answer d
The paths followed by the α particles match the predictions from (a), (b), and (c)
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
2.1.1: Practice Problems- A History of Atomic Theory (Optional) is shared under a not declared license and was authored, remixed, and/or
curated by LibreTexts.
2.1.1.5 https://chem.libretexts.org/@go/page/217249
2.2: The Structure of the Atom and How We Represent It
Skills to Develop
Describe the three subatomic particles that compose atoms
Define isotopes and give examples for several elements
Write and interpret symbols that depict the atomic number, mass number, and charge of an atom or ion
Figure 2.2.1 : If an atom could be expanded to the size of a football stadium, the nucleus would be the size of a single blueberry.
(credit middle: modification of work by “babyknight”/Wikimedia Commons; credit right: modification of work by Paxson
Woelber).
Atoms—and the protons, neutrons, and electrons that compose them—are extremely small. For example, a carbon atom weighs
less than 2 × 10−23 g, and an electron has a charge of less than 2 × 10−19 C (coulomb). When describing the properties of tiny
objects such as atoms, we use appropriately small units of measure, such as the atomic mass unit (amu) and the fundamental unit of
charge (e). The amu was originally defined based on hydrogen, the lightest element, then later in terms of oxygen. Since 1961, it
has been defined with regard to the most abundant isotope of carbon, atoms of which are assigned masses of exactly 12 amu. (This
isotope is known as “carbon-12” as will be discussed later in this module.) Thus, one amu is exactly 1/12 of the mass of one
carbon-12 atom: 1 amu = 1.6605 × 10−24 g. (The Dalton (Da) and the unified atomic mass unit (u) are alternative units that are
2.2.1 https://chem.libretexts.org/@go/page/217246
equivalent to the amu.) The fundamental unit of charge (also called the elementary charge) equals the magnitude of the charge of an
electron (e) with e = 1.602 × 10−19 C.
A proton has a mass of 1.0073 amu and a charge of 1+. A neutron is a slightly heavier particle with a mass 1.0087 amu and a
charge of zero; as its name suggests, it is neutral. The electron has a charge of 1− and is a much lighter particle with a mass of
about 0.00055 amu (it would take about 1800 electrons to equal the mass of one proton. The properties of these fundamental
particles are summarized in Table 2.2.1. (An observant student might notice that the sum of an atom’s subatomic particles does not
equal the atom’s actual mass: The total mass of six protons, six neutrons, and six electrons is 12.0993 amu, slightly larger than the
12.00 amu of an actual carbon-12 atom. This “missing” mass is known as the mass defect, and you will learn about it in the chapter
on nuclear chemistry.)
Table 2.2.1: Properties of Subatomic Particles
Name Location Charge (C) Unit Charge Mass (amu) Mass (g)
The number of protons in the nucleus of an atom is its atomic number (Z). This is the defining trait of an element: Its value
determines the identity of the atom. For example, any atom that contains six protons is the element carbon and has the atomic
number 6, regardless of how many neutrons or electrons it may have. A neutral atom must contain the same number of positive and
negative charges, so the number of protons equals the number of electrons. Therefore, the atomic number also indicates the number
of electrons in an atom. The total number of protons and neutrons in an atom is called its mass number (A). The number of
neutrons is therefore the difference between the mass number and the atomic number: A – Z = number of neutrons.
Atoms are electrically neutral if they contain the same number of positively charged protons and negatively charged electrons.
When the numbers of these subatomic particles are not equal, the atom is electrically charged and is called an ion. The charge of an
atom is defined as follows:
Atomic charge = number of protons − number of electrons
As will be discussed in more detail later in this chapter, atoms (and molecules) typically acquire charge by gaining or losing
electrons. An atom that gains one or more electrons will exhibit a negative charge and is called an anion. Positively charged atoms
called cations are formed when an atom loses one or more electrons. For example, a neutral sodium atom (Z = 11) has 11 electrons.
If this atom loses one electron, it will become a cation with a 1+ charge (11 − 10 = 1+). A neutral oxygen atom (Z = 8) has eight
electrons, and if it gains two electrons it will become an anion with a 2− charge (8 − 10 = 2−).
2.2.2 https://chem.libretexts.org/@go/page/217246
Figure 2.2.2 : (a) Insufficient iodine in the diet can cause an enlargement of the thyroid gland called a goiter. (b) The addition
of small amounts of iodine to salt, which prevents the formation of goiters, has helped eliminate this concern in the US where
salt consumption is high. (credit a: modification of work by “Almazi”/Wikimedia Commons; credit b: modification of work by
Mike Mozart)
The addition of small amounts of iodine to table salt (iodized salt) has essentially eliminated this health concern in the United
States, but as much as 40% of the world’s population is still at risk of iodine deficiency. The iodine atoms are added as anions,
and each has a 1− charge and a mass number of 127. Determine the numbers of protons, neutrons, and electrons in one of these
iodine anions.
Solution
The atomic number of iodine (53) tells us that a neutral iodine atom contains 53 protons in its nucleus and 53 electrons outside
its nucleus. Because the sum of the numbers of protons and neutrons equals the mass number, 127, the number of neutrons is 74
(127 − 53 = 74). Since the iodine is added as a 1− anion, the number of electrons is 54 [53 – (1–) = 54].
Exercise 2.2.1
An ion of platinum has a mass number of 195 and contains 74 electrons. How many protons and neutrons does it contain, and
what is its charge?
Answer
78 protons; 117 neutrons; charge is 4+
Chemical Symbols
A chemical symbol is an abbreviation that we use to indicate an element or an atom of an element. For example, the symbol for
mercury is Hg (Figure 2.2.3). We use the same symbol to indicate one atom of mercury (microscopic domain) or to label a
container of many atoms of the element mercury (macroscopic domain).
2.2.3 https://chem.libretexts.org/@go/page/217246
Figure 2.2.3 : The symbol Hg represents the element mercury regardless of the amount; it could represent one atom of mercury or
a large amount of mercury. Image used with permission from Wikipedia (user: Materialscientist).
The symbols for several common elements and their atoms are listed in Table 2.2.2. Some symbols are derived from the common
name of the element; others are abbreviations of the name in another language. Symbols have one or two letters, for example, H for
hydrogen and Cl for chlorine. To avoid confusion with other notations, only the first letter of a symbol is capitalized. For example,
Co is the symbol for the element cobalt, but CO is the notation for the compound carbon monoxide, which contains atoms of the
elements carbon (C) and oxygen (O). All known elements and their symbols are in the periodic table.
Table 2.2.2: Some Common Elements and Their Symbols
Element Symbol Element Symbol
calcium Ca magnesium Mg
chlorine Cl nitrogen N
chromium Cr oxygen O
helium He sulfur S
iodine I zinc Zn
Traditionally, the discoverer (or discoverers) of a new element names the element. However, until the name is recognized by the
International Union of Pure and Applied Chemistry (IUPAC), the recommended name of the new element is based on the Latin
word(s) for its atomic number. For example, element 106 was called unnilhexium (Unh), element 107 was called unnilseptium
(Uns), and element 108 was called unniloctium (Uno) for several years. These elements are now named after scientists or locations;
for example, element 106 is now known as seaborgium (Sg) in honor of Glenn Seaborg, a Nobel Prize winner who was active in
the discovery of several heavy elements.
IUPAC
Visit this site to learn more about IUPAC, the International Union of Pure and Applied Chemistry, and explore its periodic table.
2.2.4 https://chem.libretexts.org/@go/page/217246
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Fuse School, Open Educational Resource free of charge, under a Creative Commons License: Attribution-NonCommercial CC
BY-NC (View License Deed: https://creativecommons.org/licenses/by-nc/4.0/)
2.2: The Structure of the Atom and How We Represent It is shared under a not declared license and was authored, remixed, and/or curated by
LibreTexts.
2.2.5 https://chem.libretexts.org/@go/page/217246
2.2.1: Practice Problems- The Structure of the Atom and How We Represent It
(Optional)
PROBLEM 2.2.1.1
In what way are isotopes of a given element always different? In what way(s) are they always the same?
Answer
They always have different masses due to different numbers of neutrons.
They always have the same number of protons (which determines the identity).
PROBLEM 2.2.1.2
Write the symbol for each of the following ions:
(a) the ion with a 1+ charge, atomic number 55, and mass number 133
(b) the ion with 54 electrons, 53 protons, and 74 neutrons
(c) the ion with atomic number 15, mass number 31, and a 3− charge
(d) the ion with 24 electrons, 30 neutrons, and a 3+ charge
Answer a
133
Cs
+
55
Answer b
127
I
-
53
Answer c
31 3-
P
15
Answer d
57
Co
3+
27
PROBLEM 2.2.1.3
Write the symbol for each of the following ions:
(a) the ion with a 3+ charge, 28 electrons, and a mass number of 71
(b) the ion with 36 electrons, 35 protons, and 45 neutrons
(c) the ion with 86 electrons, 142 neutrons, and a 4+ charge
(d) the ion with a 2+ charge, atomic number 38, and mass number 87
Answer a
71 3+
31
Ga
Answer b
80 -
35
Br
Answer c
2.2.1.1 https://chem.libretexts.org/@go/page/217250
232 4+
90
Th
Answer d
87 2+
38
Sr
Problem 2.2.3
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
2.2.1: Practice Problems- The Structure of the Atom and How We Represent It (Optional) is shared under a not declared license and was authored,
remixed, and/or curated by LibreTexts.
2.2.1.2 https://chem.libretexts.org/@go/page/217250
2.3: Calculating Atomic Masses
Skills to Develop
Define the atomic mass unit and average atomic mass
Calculate average atomic mass and isotopic abundance
Define the amount unit mole and the related quantity Avogadro’s number
Explain the relation between mass, moles, and numbers of atoms or molecules, and perform calculations deriving these
quantities from one another
Video 2.3.1 : A review of counting subatomic particles and a preview of isotopes and relative atomic mass.
Isotopes
The symbol for a specific isotope of any element is written by placing the mass number as a superscript to the left of the element
symbol (Figure 2.3.4). The atomic number is sometimes written as a subscript preceding the symbol, but since this number defines
the element’s identity, as does its symbol, it is often omitted. For example, magnesium exists as a mixture of three isotopes, each
with an atomic number of 12 and with mass numbers of 24, 25, and 26, respectively. These isotopes can be identified as 24Mg,
25
Mg, and 26Mg. These isotope symbols are read as “element, mass number” and can be symbolized consistent with this reading.
For instance, 24Mg is read as “magnesium 24,” and can be written as “magnesium-24” or “Mg-24.” 25Mg is read as “magnesium
25,” and can be written as “magnesium-25” or “Mg-25.” All magnesium atoms have 12 protons in their nucleus. They differ only
because a 24Mg atom has 12 neutrons in its nucleus, a 25Mg atom has 13 neutrons, and a 26Mg has 14 neutrons.
Figure 2.3.1 : The symbol for an atom indicates the element via its usual two-letter symbol, the mass number as a left superscript,
the atomic number as a left subscript (sometimes omitted), and the charge as a right superscript.
Information about the naturally occurring isotopes of elements with atomic numbers 1 through 10 is given in Table 2.3.2. Note that
in addition to standard names and symbols, the isotopes of hydrogen are often referred to using common names and accompanying
symbols. Hydrogen-2, symbolized 2H, is also called deuterium and sometimes symbolized D. Hydrogen-3, symbolized 3H, is also
called tritium and sometimes symbolized T.
Table 2.3.1: Nuclear Compositions of Atoms of the Very Light Elements
Number of Number of % Natural
Element Symbol Atomic Number Mass (amu)
Protons Neutrons Abundance
2.3.1 https://chem.libretexts.org/@go/page/217247
Number of Number of % Natural
Element Symbol Atomic Number Mass (amu)
Protons Neutrons Abundance
1
H
1
1 1 0 1.0078 99.989
(protium)
2
H
hydrogen 1
1 1 1 2.0141 0.0115
(deuterium)
3
H
1
1 1 2 3.01605 — (trace)
(tritium)
3
2
He 2 2 1 3.01603 0.00013
helium
4
2
He 2 2 2 4.0026 100
6
3
Li 3 3 3 6.0151 7.59
lithium
7
3
Li 3 3 4 7.0160 92.41
beryllium 9
4
Be 4 4 5 9.0122 100
10
5
B 5 5 5 10.0129 19.9
boron
11
5
B 5 5 6 11.0093 80.1
12
6
C 6 6 6 12.0000 98.89
carbon 13
6
C 6 6 7 13.0034 1.11
14
6
C 6 6 8 14.0032 — (trace)
14
7
N 7 7 7 14.0031 99.63
nitrogen
15
7
N 7 7 8 15.0001 0.37
16
8
O 8 8 8 15.9949 99.757
oxygen 17
8
O 8 8 9 16.9991 0.038
18
8
O 8 8 10 17.9992 0.205
fluorine 19
9
F 9 9 10 18.9984 100
20
10
Ne 10 10 10 19.9924 90.48
neon 21
10
Ne 10 10 11 20.9938 0.27
22
10
Ne 10 10 12 21.9914 9.25
2.3.2 https://chem.libretexts.org/@go/page/217247
Build an Atom
Symbol
Atom
Use this Build an Atom simulator to build atoms of the first 10 elements, see which isotopes exist, check nuclear stability, and gain
experience with isotope symbols.
Atomic Mass
Because each proton and each neutron contribute approximately one amu to the mass of an atom, and each electron contributes far
less, the atomic mass of a single atom is approximately equal to its mass number (a whole number). However, the average masses
of atoms of most elements are not whole numbers because most elements exist naturally as mixtures of two or more isotopes.
The mass of an element shown in a periodic table or listed in a table of atomic masses is a weighted, average mass of all the
isotopes present in a naturally occurring sample of that element. This is equal to the sum of each individual isotope’s mass
multiplied by its fractional abundance.
For example, the element boron is composed of two isotopes: About 19.9% of all boron atoms are 10B with a mass of 10.0129 amu,
and the remaining 80.1% are 11B with a mass of 11.0093 amu. The average atomic mass for boron is calculated to be:
2.3.3 https://chem.libretexts.org/@go/page/217247
boron average mass = (0.199 × 10.0129 amu) + (0.801 × 11.0093 amu)
= 10.81 amu
It is important to understand that no single boron atom weighs exactly 10.8 amu; 10.8 amu is the average mass of all boron atoms,
and individual boron atoms weigh either approximately 10 amu or 11 amu.
average mass = (0.9184 × 19.9924 amu) + (0.0047 × 20.9940 amu) + (0.0769 × 21.9914 amu)
= 20.15 amu
The average mass of a neon atom in the solar wind is 20.15 amu. (The average mass of a terrestrial neon atom is 20.1796 amu.
This result demonstrates that we may find slight differences in the natural abundance of isotopes, depending on their origin.)
Exercise 2.3.1
A sample of magnesium is found to contain 78.70% of 24Mg atoms (mass 23.98 amu), 10.13% of 25
Mg atoms (mass 24.99
amu), and 11.17% of 26Mg atoms (mass 25.98 amu). Calculate the average mass of a Mg atom.
Answer
24.31 amu
We can also do variations of this type of calculation, as shown in the next example.
If we let x represent the fraction that is 35Cl, then the fraction that is 37Cl is represented by 1.00 − x.
35 37 37
(The fraction that is Cl + the fraction that is Cl must add up to 1, so the fraction of Cl must equal 1.00 − the fraction of
35
Cl.)
Substituting this into the average mass equation, we have:
1.99705x = 1.513
1.513
x = = 0.7576
1.99705
So solving yields: x = 0.7576, which means that 1.00 − 0.7576 = 0.2424. Therefore, chlorine consists of 75.76% 35Cl and
24.24% 37Cl.
Exercise 2.3.2
2.3.4 https://chem.libretexts.org/@go/page/217247
Naturally occurring copper consists of 63Cu (mass 62.9296 amu) and 65Cu (mass 64.9278 amu), with an average mass of 63.546
amu. What is the percent composition of Cu in terms of these two isotopes?
Answer
69.15% Cu-63 and 30.85% Cu-65
Mixtures
Isotopes
Summary
2.3.5 https://chem.libretexts.org/@go/page/217247
The Nucleus: Crash Course Chemistry #1
Video 2.3.6 : Watch this video for a review of relative atomic mass and isotopes.
An atom consists of a small, positively charged nucleus surrounded by electrons. The nucleus contains protons and neutrons; its
diameter is about 100,000 times smaller than that of the atom. The mass of one atom is usually expressed in atomic mass units
(amu), which is referred to as the atomic mass. An amu is defined as exactly 1/12 of the mass of a carbon-12 atom and is equal to
1.6605 × 10−24 g.
Protons are relatively heavy particles with a charge of 1+ and a mass of 1.0073 amu. Neutrons are relatively heavy particles with no
charge and a mass of 1.0087 amu. Electrons are light particles with a charge of 1− and a mass of 0.00055 amu. The number of
protons in the nucleus is called the atomic number (Z) and is the property that defines an atom’s elemental identity. The sum of the
numbers of protons and neutrons in the nucleus is called the mass number and, expressed in amu, is approximately equal to the
mass of the atom. An atom is neutral when it contains equal numbers of electrons and protons.
Isotopes of an element are atoms with the same atomic number but different mass numbers; isotopes of an element, therefore, differ
from each other only in the number of neutrons within the nucleus. When a naturally occurring element is composed of several
isotopes, the atomic mass of the element represents the average of the masses of the isotopes involved. A chemical symbol
identifies the atoms in a substance using symbols, which are one-, two-, or three-letter abbreviations for the atoms.
Looking Beyond
Video 2.3.7 : Remember our exploration into the size of an atom last week? This video goes deeper into investigating the size of
the subatomic particles we just discussed.
Footnotes
1. Read more about the redefinition of SI units including the kilogram here (Laura Howe, CE&N, Nov. 16, 2018).
2.3.6 https://chem.libretexts.org/@go/page/217247
Key Equations
average mass = ∑ (f ractional abundance × isotopic mass)i
i
Glossary
anion
negatively charged atom or molecule (contains more electrons than protons)
atomic mass
average mass of atoms of an element, expressed in amu
cation
positively charged atom or molecule (contains fewer electrons than protons)
chemical symbol
one-, two-, or three-letter abbreviation used to represent an element or its atoms
Dalton (Da)
alternative unit equivalent to the atomic mass unit
ion
electrically charged atom or molecule (contains unequal numbers of protons and electrons)
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
2.3.7 https://chem.libretexts.org/@go/page/217247
TED-Ed’s commitment to creating lessons worth sharing is an extension of TED’s mission of spreading great ideas. Within
TED-Ed’s growing library of TED-Ed animations, you will find carefully curated educational videos, many of which represent
collaborations between talented educators and animators nominated through the TED-Ed website.
2.3: Calculating Atomic Masses is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
2.3.8 https://chem.libretexts.org/@go/page/217247
2.3.1: Practice Problems- Calculating Atomic Masses (Optional)
PROBLEM 2.3.1.1
Determine the number of protons, neutrons, and electrons in the following isotopes that are used in medical diagnoses:
(a) atomic number 9, mass number 18, charge of 1−
(b) atomic number 43, mass number 99, charge of 7+
(c) atomic number 53, atomic mass number 131, charge of 1−
(d) atomic number 81, atomic mass number 201, charge of 1+
(e) Name the elements in parts (a), (b), (c), and (d)
Answer a
p: 9; n: 9; e: 10
Answer b
p: 43; n: 56; e: 36
Answer c
p: 53; n: 78; e: 54
Answer d
p: 81; n: 120; e: 80
Answer e
a - F; b - Tc; c - I; d - Tl
Problem 2.3.1
PROBLEM 2.3.1.2
Give the number of protons, electrons, and neutrons in neutral atoms of each of the following isotopes:
(a) 10
5
B
(b) 199
80
Hg
(c) 63
29
Cu
(d) 13
6
C
2.3.1.1 https://chem.libretexts.org/@go/page/217251
(e) 77
34
Se
Answer a
p&e: 5; n: 5
Answer b
p&e: 80; n: 119
Answer c
p&e: 29; n: 34
Answer d
p&e: 6; n: 7
Answer e
p&e: 34; n: 43
PROBLEM 2.3.1.3
An element has the following natural abundances and isotopic masses: 90.92% abundance with 19.99 amu, 0.26% abundance
with 20.99 amu, and 8.82% abundance with 21.99 amu. Calculate the average atomic mass of this element.
Answer
20.16 amu
Problem 2.3.3
PROBLEM 2.3.1.4
Average atomic masses listed by IUPAC are based on a study of experimental results. Bromine has two isotopes, 79Br and 81Br,
whose masses (78.9183 and 80.9163 amu) and abundances (50.69% and 49.31%) were determined in earlier experiments.
Calculate the average atomic mass of Br based on these experiments. How does this compare to the value given on the periodic
table?
Answer
79.90 amu; this matches the value on the periodic table
PROBLEM 2.3.1.5
The 18O:16O abundance ratio in some meteorites is greater than that used to calculate the average atomic mass of oxygen on
earth. Is the average atomic mass of an oxygen atom in these meteorites greater than, less than, or equal to a terrestrial oxygen
2.3.1.2 https://chem.libretexts.org/@go/page/217251
atom?
Answer
18
Greater, since the contribution to the average atomic mass of O is greater, that will raise the average atomic mass in
meteorites compared to on earth.
PROBLEM 2.3.1.6
Compare 1 mole of H2, 1 mole of O2, and 1 mole of F2.
(a) Which has the largest number of molecules? Explain why.
(b) Which has the greatest mass? Explain why.
Answer a
1 mole is always 6.022 x 1023 molecules. They have the same number of molecules.
Answer b
F2; it has the highest molar mass.
PROBLEM 2.3.1.7
Which contains the greatest mass of oxygen: 0.75 mol of ethanol (C2H5OH), 0.60 mol of formic acid (HCO2H), or 1.0 mol of
water (H2O)? Explain why.
Answer
Formic acid. Its formula has twice as many oxygen atoms as the other two compounds (one each). Therefore, 0.60 mol of
formic acid would be equivalent to 1.20 mol of a compound containing a single oxygen atom.
PROBLEM 2.3.1.8
Determine the mass of each of the following:
(a) 0.0146 mol KOH
(b) 10.2 mol ethane, C2H6
(c) 1.6 × 10−3 mol Na2SO4
(d) 6.854 × 103 mol glucose, C6H12O6
(e) 2.86 mol Co(NH3)6Cl3
Answer a
0.819 g
Answer b
307 g
Answer c
0.23 g
Answer d
1.235 × 106 g (1235 kg)
Answer e
765 g
PROBLEM 2.3.1.9
2.3.1.3 https://chem.libretexts.org/@go/page/217251
Which of the following represents the least number of molecules?
a. 20.0 g of H2O (18.02 g/mol)
b. 77.0 g of CH4 (16.06 g/mol)
c. 68.0 g of CaH2 (42.09 g/mol)
d. 100.0 g of N2O (44.02 g/mol)
e. 84.0 g of HF (20.01 g/mol)
Answer
20.0 g of H2O represents the smallest number of moles, meaning the least number of molecules present. Since 1 mole =
6.022 × 1023 molecules (or atoms) regardless of identity, the least number of moles will equal the least number of molecules.
Problem 2.3.9
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
2.3.1: Practice Problems- Calculating Atomic Masses (Optional) is shared under a not declared license and was authored, remixed, and/or curated
by LibreTexts.
2.3.1.4 https://chem.libretexts.org/@go/page/217251
2.4: Using the Mole in Calculations
The Mole
The identity of a substance is defined not only by the types of atoms or ions it contains, but by the quantity of each type of atom or
ion. For example, water, H2O, and hydrogen peroxide, H2O2, are alike in that their respective molecules are composed of hydrogen
and oxygen atoms. However, because a hydrogen peroxide molecule contains two oxygen atoms, as opposed to the water molecule,
which has only one, the two substances exhibit very different properties. Today, we possess sophisticated instruments that allow the
direct measurement of these defining microscopic traits; however, the same traits were originally derived from the measurement of
macroscopic properties (the masses and volumes of bulk quantities of matter) using relatively simple tools (balances and
volumetric glassware). This experimental approach required the introduction of a new unit for amount of substances, the mole,
which remains indispensable in modern chemical science.
The mole is an amount unit similar to familiar units like pair, dozen, gross, etc. It provides a specific measure of the number of
atoms or molecules in a bulk sample of matter. A mole is defined as the amount of substance containing the same number of
discrete entities (such as atoms, molecules, and ions) as the number of atoms in a sample of pure 12C weighing exactly 12 g. One
Latin connotation for the word “mole” is “large mass” or “bulk,” which is consistent with its use as the name for this unit. The
mole provides a link between an easily measured macroscopic property, bulk mass, and an extremely important fundamental
property, number of atoms, molecules, and so forth.
The number of entities composing a mole has been experimentally determined to be 6.02214179 × 10 , a fundamental constant
23
named Avogadro’s number (NA) or the Avogadro constant in honor of Italian scientist Amedeo Avogadro. This constant is properly
reported with an explicit unit of “per mole,” a conveniently rounded version being 6.022 \times 10^{23}\/ce{mol}\).
Consistent with its definition as an amount unit, 1 mole of any element contains the same number of atoms as 1 mole of any other
element. The masses of 1 mole of different elements, however, are different, since the masses of the individual atoms are drastically
different. The molar mass of an element (or compound) is the mass in grams of 1 mole of that substance, a property expressed in
units of grams per mole (g/mol) (Figure 2.4.1).
2.4.1 https://chem.libretexts.org/@go/page/218823
Figure 2.4.1 : Each sample contains 6.022 × 10 atoms —1.00 mol of atoms. From left to right (top row): 65.4 g zinc, 12.0 g
23
carbon, 24.3 g magnesium, and 63.5 g copper. From left to right (bottom row): 32.1 g sulfur, 28.1 g silicon, 207 g lead, and 118.7 g
tin. (credit: modification of work by Mark Ott).
Because the definitions of both the mole and the atomic mass unit are based on the same reference substance, 12C, the molar mass
of any substance is numerically equivalent to its atomic or formula weight in amu. Per the amu definition, a single 12C atom weighs
12 amu (its atomic mass is 12 amu). The former definition of the mole was that a mole was 12 g of 12C contains 1 mole of 12C
atoms (its molar mass is 12 g/mol). This relationship holds for all elements, since their atomic masses are measured relative to that
of the amu-reference substance, 12C. Extending this principle, the molar mass of a compound in grams is likewise numerically
equivalent to its formula mass in amu. On May 20, 2019 the definition was permanently changed to Avogadro's number: a mole is
6.02214179 × 10
23
of any object, from atoms to apples.1
While atomic mass and molar mass are numerically equivalent, keep in mind that they are vastly different in terms of scale, as
represented by the vast difference in the magnitudes of their respective units (amu versus g). To appreciate the enormity of the
mole, consider a small drop of water after a rainfall. Although this represents just a tiny fraction of 1 mole of water (~18 g), it
contains more water molecules than can be clearly imagined. If the molecules were distributed equally among the roughly seven
billion people on earth, each person would receive more than 100 billion molecules.
2.4.2 https://chem.libretexts.org/@go/page/218823
How big is a mole? (Not the animal, the …
Video 2.4.2 : The mole is used in chemistry to represent 6.022 × 10 of something, but it can be difficult to conceptualize such a
23
large number. Watch this video and then complete the “Think” questions that follow. Explore more about the mole by reviewing the
information under “Dig Deeper.”
The relationships between formula mass, the mole, and Avogadro’s number can be applied to compute various quantities that
describe the composition of substances and compounds. For example, if we know the mass and chemical composition of a
substance, we can determine the number of moles and calculate number of atoms or molecules in the sample. Likewise, if we know
the number of moles of a substance, we can derive the number of atoms or molecules and calculate the substance’s mass.
The factor-label method supports this mathematical approach since the unit “g” cancels and the answer has units of “mol:”
mol K
4.7 g K( ) = 0.12 mol K
39.10 g
The calculated magnitude (0.12 mol K) is consistent with our ballpark expectation, since it is a bit greater than 0.1 mol.
Answer
0.360 mol
2.4.3 https://chem.libretexts.org/@go/page/218823
Solution
The molar amount of Ar is provided and must be used to derive the corresponding mass in grams. Since the amount of Ar is less
than 1 mole, the mass will be less than the mass of 1 mole of Ar, approximately 40 g. The molar amount in question is
approximately one-one thousandth (~10−3) of a mole, and so the corresponding mass should be roughly one-one thousandth of
the molar mass (~0.04 g):
In this case, logic dictates (and the factor-label method supports) multiplying the provided amount (mol) by the molar mass
(g/mol):
−4
39.95 g
9.2 × 10 mol Ar ( ) = 0.037 g Ar
mol Ar
Exercise 2.4.2
What is the mass of 2.561 mol of gold?
Answer
504.4 g
Figure 2.4.6 : Copper wire is composed of many, many atoms of Cu. (credit: Emilian Robert Vicol)
Solution
The number of Cu atoms in the wire may be conveniently derived from its mass by a two-step computation: first calculating the
molar amount of Cu, and then using Avogadro’s number (NA) to convert this molar amount to number of Cu atoms:
Considering that the provided sample mass (5.00 g) is a little less than one-tenth the mass of 1 mole of Cu (~64 g), a reasonable
estimate for the number of atoms in the sample would be on the order of one-tenth NA, or approximately 1022 Cu atoms.
Carrying out the two-step computation yields:
23
mol Cu 6.022 × 10 atoms 22
5.00 g Cu ( )( ) = 4.74 × 10 atoms of copper (2.4.1)
63.55 g
mol
2.4.4 https://chem.libretexts.org/@go/page/218823
The factor-label method yields the desired cancellation of units, and the computed result is on the order of 1022 as expected.
Exercise 2.4.3
A prospector panning for gold in a river collects 15.00 g of pure gold. How many Au atoms are in this quantity of gold?
Answer
4.586 × 10
22
Au atoms
The molar mass of glycine is required for this calculation, and it is computed in the same fashion as its molecular mass. One
mole of glycine, C2H5O2N, contains 2 moles of carbon, 5 moles of hydrogen, 2 moles of oxygen, and 1 mole of nitrogen:
The provided mass of glycine (~28 g) is a bit more than one-third the molar mass (~75 g/mol), so we would expect the
computed result to be a bit greater than one-third of a mole (~0.33 mol). Dividing the compound’s mass by its molar mass
yields:
mol glycine
28.35 g glycine ( ) = 0.378 mol glycine
75.07 g
Exercise 2.4.4
How many moles of sucrose, C 12 H22 O11 , are in a 25-g sample of sucrose?
Answer
0.073 mol
2.4.5 https://chem.libretexts.org/@go/page/218823
The molar mass for this compound is computed to be 176.124 g/mol. The given number of moles is a very small fraction of a
mole (~10−4 or one-ten thousandth); therefore, we would expect the corresponding mass to be about one-ten thousandth of the
molar mass (~0.02 g). Performing the calculation, we get:
176.124 g
−4
1.42 × 10 mol vitamin C ( ) = 0.0250 g vitamin C
mol vitamin C
Exercise 2.4.5
What is the mass of 0.443 mol of hydrazine, N 2 H4 ?
Answer
14.2 g
Example 2.4.6 : Deriving the Number of Molecules from the Compound Mass
A packet of an artificial sweetener contains 40.0 mg of saccharin (C7H5NO3S), which has the structural formula:
Given that saccharin has a molar mass of 183.18 g/mol, how many saccharin molecules are in a 40.0-mg (0.0400-g) sample of
saccharin? How many carbon atoms are in the same sample?
Solution
The number of molecules in a given mass of compound is computed by first deriving the number of moles, as demonstrated in
Example 2.4.8, and then multiplying by Avogadro’s number:
Using the provided mass and molar mass for saccharin yields:
23
mol C H NO S 6.022 × 10 C H NO S molecules
7 5 3 7 5 3
0.0400 g C H NO S ( )( ) (2.4.2)
7 5 3
183.18 g C H NO S 1 mol C H NO S
7 5 3
7 5 3
20
= 1.31 × 10 C H NO S molecules
7 5 3
The compound’s formula shows that each molecule contains seven carbon atoms, and so the number of C atoms in the provided
sample is:
7 C atoms
20 21
1.31 × 10 C H NO S molecules ( ) = 9.20 × 10 C atoms
7 5 3
1 C H NO S molecule
7 5 3
2.4.6 https://chem.libretexts.org/@go/page/218823
Exercise 2.4.6
How many C 4 H10 molecules are contained in 9.213 g of this compound? How many hydrogen atoms?
Answer
22
9.545 × 10 molecules C4 H10
23
9.545 × 10 atoms H
Summary
Video 2.4.5 : A preview of some of the uses we will have for moles in upcoming units
Glossary
mole
amount of substance containing the same number of atoms, molecules, ions, or other entities as the number of atoms in exactly
12 grams of 12C
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
2.4: Using the Mole in Calculations is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
2.4.7 https://chem.libretexts.org/@go/page/218823
2.5: Highlighting Instrumentation- Mass Spectrometry
Skills to Develop
Define the atomic mass unit and average atomic mass
Calculate average atomic mass and isotopic abundance
Define the amount unit mole and the related quantity Avogadro’s number
Explain the relation between mass, moles, and numbers of atoms or molecules, and perform calculations deriving these
quantities from one another
Mass Spectrometry
The occurrence and natural abundances of isotopes can be experimentally determined using an instrument called a mass
spectrometer. Mass spectrometry (MS) is widely used in chemistry, forensics, medicine, environmental science, and many other
fields to analyze and help identify the substances in a sample of material. In a typical mass spectrometer (Figure 2.5.1), the sample
is vaporized and exposed to a high-energy electron beam that causes the sample’s atoms (or molecules) to become electrically
charged, typically by losing one or more electrons. These cations then pass through a (variable) electric or magnetic field that
deflects each cation’s path to an extent that depends on both its mass and charge (similar to how the path of a large steel ball
bearing rolling past a magnet is deflected to a lesser extent that that of a small steel BB). The ions are detected, and a plot of the
relative number of ions generated versus their mass-to-charge ratios (a mass spectrum) is made. The height of each vertical feature
or peak in a mass spectrum is proportional to the fraction of cations with the specified mass-to-charge ratio. Since its initial use
during the development of modern atomic theory, MS has evolved to become a powerful tool for chemical analysis in a wide range
of applications.
Figure 2.5.1 : Analysis of zirconium in a mass spectrometer produces a mass spectrum with peaks showing the different isotopes
of Zr.
T
2.5.1 https://chem.libretexts.org/@go/page/217248
Mass Spectrometry MS
Video 2.5.1 : Watch this video from the Royal Society for Chemistry for a brief description of the rudiments of mass spectrometry.
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
2.5: Highlighting Instrumentation- Mass Spectrometry is shared under a not declared license and was authored, remixed, and/or curated by
LibreTexts.
2.5.2 https://chem.libretexts.org/@go/page/217248
CHAPTER OVERVIEW
3: Nuclei, Ions and the Periodic Table
3.1: Nuclear Chemistry and Radioactive Decay
3.1.1: Practice Problems- Nuclear Chemistry and Radioactive Decay (Optional)
3.2: Organization of the Periodic Table
3.2.1: Practice Problems - Organization of the Periodic Table (Optional)
3.3: Predicting Charges of Ions
3: Nuclei, Ions and the Periodic Table is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
1
3.1: Nuclear Chemistry and Radioactive Decay
Skills to Develop
Describe nuclear structure in terms of protons, neutrons, and electrons
Recognize the role of mass defect and binding energy for nuclei in atomic masses
Explain trends in the relative stability of nuclei
Identify common particles and energies involved in nuclear reactions
Write and balance nuclear equations
Recognize common modes of radioactive decay
Describe kinetic parameters for decay processes, including half-life
Describe common radiometric dating techniques
Video 3.1.1 : A video review of isotopes (Unit 2), with a peek at radioactive isotopes.
neutrons. Recall that the number of protons in the nucleus is called the atomic number (Z ) of the element, and the sum of the
number of protons and the number of neutrons is the mass number (A ). Atoms with the same atomic number but different mass
numbers are isotopes of the same element. When referring to a single type of nucleus, we often use the term nuclide and identify it
by the notation:
A
X (3.1.1)
Z
where
X is the symbol for the element,
A is the mass number, and
Z is the atomic number.
Often a nuclide is referenced by the name of the element followed by a hyphen and the mass number. For example, 14
6
C is called
“carbon-14.”
Protons and neutrons, collectively called nucleons, are packed together tightly in a nucleus. With a radius of about 10−15 meters, a
nucleus is quite small compared to the radius of the entire atom, which is about 10−10 meters. Nuclei are extremely dense compared
to bulk matter, averaging 1.8 × 10 grams per cubic centimeter. For example, water has a density of 1 gram per cubic centimeter,
14
and iridium, one of the densest elements known, has a density of 22.6 g/cm3. If the earth’s density were equal to the average
nuclear density, the earth’s radius would be only about 200 meters (earth’s actual radius is approximately 6.4 × 10 meters, 30,000
6
times larger).
Recall our exploration of atomic size with this video:
3.1.1 https://chem.libretexts.org/@go/page/217253
Just How Small is an Atom?
However, mass spectrometric measurements reveal that the mass of an He atom is 4.0026 amu, less than the combined masses of
4
2
its six constituent subatomic particles. This difference between the calculated and experimentally measured masses is known as the
mass defect of the atom. In the case of helium, the mass defect indicates a “loss” in mass of 4.0331 amu – 4.0026 amu = 0.0305
amu. The loss in mass accompanying the formation of an atom from protons, neutrons, and electrons is due to the conversion of
that mass into energy that is evolved as the atom forms. The nuclear binding energy is the energy produced when the atoms’
nucleons are bound together; this is also the energy needed to break a nucleus into its constituent protons and neutrons. In
comparison to chemical bond energies, nuclear binding energies are vastly greater, as we will learn in this section. Consequently,
the energy changes associated with nuclear reactions are vastly greater than are those for chemical reactions.
The conversion between mass and energy is most identifiably represented by the mass-energy equivalence equation as stated by
Albert Einstein:
2
E = mc (3.1.3)
where E is energy, m is mass of the matter being converted, and c is the speed of light in a vacuum. This equation can be used to
find the amount of energy that results when matter is converted into energy. Using this mass-energy equivalence equation, the
nuclear binding energy of a nucleus may be calculated from its mass defect, a calculation beyond the scope of our course. A
variety of units are commonly used for nuclear binding energies, including electron volts (eV), with 1 eV equaling the amount of
energy necessary to the move the charge of an electron across an electric potential difference of 1 volt, making
J.
−19
1 eV = 1.602 × 10
Because the energy changes for breaking and forming bonds are so small compared to the energy changes for breaking or forming
nuclei, the changes in mass during all ordinary chemical reactions are virtually undetectable. As we will discuss later in our unit on
thermochemistry, the most energetic chemical reactions exhibit enthalpies on the order of thousands of kJ/mol, which is equivalent
3.1.2 https://chem.libretexts.org/@go/page/217253
to mass differences in the nanogram range (10–9 g). On the other hand, nuclear binding energies are typically on the order of
billions of kJ/mol, corresponding to mass differences in the milligram range (10–3 g).
Nuclear Stability
A nucleus is stable if it cannot be transformed into another configuration without adding energy from the outside. Of the thousands
of nuclides that exist, about 250 are stable. A plot of the number of neutrons versus the number of protons for stable nuclei reveals
that the stable isotopes fall into a narrow band. This region is known as the band of stability (also called the belt, zone, or valley of
stability). The straight line in Figure 3.1.1 represents nuclei that have a 1:1 ratio of protons to neutrons (n:p ratio). Note that the
lighter stable nuclei, in general, have equal numbers of protons and neutrons. For example, nitrogen-14 has seven protons and
seven neutrons. Heavier stable nuclei, however, have increasingly more neutrons than protons. For example: iron-56 has 30
neutrons and 26 protons, an n:p ratio of 1.15, whereas the stable nuclide lead-207 has 125 neutrons and 82 protons, an n:p ratio
equal to 1.52. This is because larger nuclei have more proton-proton repulsions, and require larger numbers of neutrons to provide
compensating strong forces to overcome these electrostatic repulsions and hold the nucleus together.
Figure 3.1.1 : This plot shows the nuclides that are known to exist and those that are stable. The stable nuclides are indicated in
blue, and the unstable nuclides are indicated in green. Note that all isotopes of elements with atomic numbers greater than 83 are
unstable. The solid line is the line where n = Z.
The nuclei that are to the left or to the right of the band of stability are unstable and exhibit radioactivity. They change
spontaneously (decay) into other nuclei that are either in, or closer to, the band of stability. These nuclear decay reactions convert
one unstable isotope (or radioisotope) into another, more stable, isotope. We will discuss the nature and products of this radioactive
decay in subsequent sections of this unit.
Several observations may be made regarding the relationship between the stability of a nucleus and its structure. Nuclei with even
numbers of protons, neutrons, or both are more likely to be stable (Table 3.1.1). Nuclei with certain numbers of nucleons, known as
magic numbers, are stable against nuclear decay. These numbers of protons or neutrons (2, 8, 20, 28, 50, 82, and 126) make
complete shells in the nucleus. These are similar in concept to the stable electron shells observed for the noble gases. Nuclei that
have magic numbers of both protons and neutrons, such as He, O, Ca, and Pb and are particularly stable. These trends in
4
2
16
8
40
20
208
82
nuclear stability may be rationalized by considering a quantum mechanical model of nuclear energy states analogous to that used to
describe electronic states, which we will discuss later in this course. The details of this model are beyond the scope of this course.
Table 3.1.1: Stable Nuclear Isotopes
3.1.3 https://chem.libretexts.org/@go/page/217253
Number of Stable Isotopes Proton Number Neutron Number
53 even odd
50 odd even
5 odd odd
The relative stability of a nucleus is correlated with its binding energy per nucleon, the total binding energy for the nucleus divided
by the number or nucleons in the nucleus. For instance, the binding energy for a He nucleus is therefore:
4
2
28.4 MeV
= 7.10 MeV/nucleon (3.1.4)
4 nucleons
The binding energy per nucleon of a nuclide on the curve shown in Figure 3.1.2
Figure 3.1.2 : The binding energy per nucleon is largest for nuclides with mass number of approximately 56.
Changes of nuclei that result in changes in their atomic numbers, mass numbers, or energy states are nuclear reactions. To
describe a nuclear reaction, we use an equation that identifies the nuclides involved in the reaction, their mass numbers and atomic
numbers, and the other particles involved in the reaction.
3.1.4 https://chem.libretexts.org/@go/page/217253
Video 3.1.3 : A brief overview of the different types of radioactivity.
Many entities can be involved in nuclear reactions. The most common are protons, neutrons, alpha particles, beta particles,
positrons, and gamma rays, as shown in Figure 3.1.3. Protons ( p, also represented by the symbol H) and neutrons ( n) are the
1
1
1
1
1
0
constituents of atomic nuclei, and have been described previously. Alpha particles ( He , also represented by the symbol α) are
4
2
4
2
high-energy helium nuclei. Beta particles ( β , also represented by the symbol e) are high-energy electrons, and gamma rays
−1
0
−1
0
are photons of very high-energy electromagnetic radiation. Positrons ( e , also represented by the symbol β) are positively
0
+1 +1
0
charged electrons (“anti-electrons”). The subscripts and superscripts are necessary for balancing nuclear equations, but are usually
optional in other circumstances. For example, an alpha particle is a helium nucleus (He) with a charge of +2 and a mass number of
4, so it is symbolized He. This works because, in general, the ion charge is not important in the balancing of nuclear equations.
4
2
Figure 3.1.3 : Although many species are encountered in nuclear reactions, this table summarizes the names, symbols,
representations, and descriptions of the most common of these.
Note that positrons are exactly like electrons, except they have the opposite charge. They are the most common example of
antimatter, particles with the same mass but the opposite state of another property (for example, charge) than ordinary matter.
When antimatter encounters ordinary matter, both are annihilated and their mass is converted into energy in the form of gamma
rays (γ)—and other much smaller subnuclear particles, which are beyond the scope of this chapter—according to the mass-energy
equivalence equation E = mc , seen in the preceding section. For example, when a positron and an electron collide, both are
2
Gamma rays compose short wavelength, high-energy electromagnetic radiation and are (much) more energetic than better-known
X-rays. Gamma rays are produced when a nucleus undergoes a transition from a higher to a lower energy state, similar to how a
photon is produced by an electronic transition from a higher to a lower energy level. Due to the much larger energy differences
between nuclear energy shells, gamma rays emanating from a nucleus have energies that are typically millions of times larger than
electromagnetic radiation emanating from electronic transitions.
3.1.5 https://chem.libretexts.org/@go/page/217253
2. The sum of the charges of the reactants equals the sum of the charges of the products.
If the atomic number and the mass number of all but one of the particles in a nuclear reaction are known, we can identify the
particle by balancing the reaction. For instance, we could determine that O is a product of the nuclear reaction of N and He if
17
8
14
7
4
2
we knew that a proton, H , was one of the two products. Example 3.1.1 shows how we can identify a nuclide by balancing the
1
1
nuclear reaction.
where
A is the mass number and
Z is the atomic number of the new nuclide, X .
Because the sum of the mass numbers of the reactants must equal the sum of the mass numbers of the products:
25 + 4 = A + 1 (3.1.7)
so
A = 28 (3.1.8)
12 + 2 = Z + 1 (3.1.9)
so
Z = 13 (3.1.10)
Check the periodic table: The element with nuclear charge = +13 is aluminum. Thus, the product is 28
13
Al .
Exercise 3.1.1
The nuclide 125
53
I combines with an electron and produces a new nucleus and no other massive particles. What is the equation for
this reaction?
Answer:
125 0 125
I+ e → Te
53 −1 52
Following are the equations of several nuclear reactions that have important roles in the history of nuclear chemistry:
The first naturally occurring unstable element that was isolated, polonium, was discovered by the Polish scientist Marie Curie
and her husband Pierre in 1898. It decays, emitting α particles:
212 208 4
Po ⟶ Pb + He (3.1.11)
84 82 2
The first nuclide to be prepared by artificial means was an isotope of oxygen, 17O. It was made by Ernest Rutherford in 1919 by
bombarding nitrogen atoms with α particles:
14 4 17 1
7
N+ α ⟶ 8
O + 1H (3.1.12)
2
James Chadwick discovered the neutron in 1932, as a previously unknown neutral particle produced along with 12C by the
nuclear reaction between 9Be and 4He:
9 4 12 1
Be + He ⟶ C+ n (3.1.13)
4 2 6 0
3.1.6 https://chem.libretexts.org/@go/page/217253
The first element to be prepared that does not occur naturally on the earth, technetium, was created by bombardment of
molybdenum by deuterons (heavy hydrogen, H ), by Emilio Segre and Carlo Perrier in 1937:
2
1
2 97 1 97
1
H + 42 Mo ⟶ 2 n+ Tc (3.1.14)
0 43
The first controlled nuclear chain reaction was carried out in a reactor at the University of Chicago in 1942. One of the many
reactions involved was:
235 1 87 146 1
92
U+ n ⟶ 35
Br + 57
La + 3 n (3.1.15)
0 0
Following the somewhat serendipitous discovery of radioactivity by Becquerel, many prominent scientists began to investigate this
new, intriguing phenomenon. Among them were Marie Curie (Video 3.1.4; the first woman to win a Nobel Prize, and the only
person to win two Nobel Prizes in different sciences—chemistry and physics), who was the first to coin the term “radioactivity,”
and Ernest Rutherford (of gold foil experiment fame), who investigated and named three of the most common types of radiation.
During the beginning of the twentieth century, many radioactive substances were discovered, the properties of radiation were
investigated and quantified, and a solid understanding of radiation and nuclear decay was developed.
Learn More about Marie Curie's Life & Work
Video 3.1.4 : A brief overview of the life and work of Marie Curie.
The spontaneous change of an unstable nuclide into another is radioactive decay. The unstable nuclide is called the parent nuclide;
the nuclide that results from the decay is known as the daughter nuclide. The daughter nuclide may be stable, or it may decay itself.
The radiation produced during radioactive decay is such that the daughter nuclide lies closer to the band of stability than the parent
nuclide, so the location of a nuclide relative to the band of stability can serve as a guide to the kind of decay it will undergo (Figure
3.1.4).
Figure 3.1.4 : A nucleus of uranium-238 (the parent nuclide) undergoes α decay to form thorium-234 (the daughter nuclide). The
alpha particle removes two protons (green) and two neutrons (gray) from the uranium-238 nucleus.
Although the radioactive decay of a nucleus is too small to see with the naked eye, we can indirectly view radioactive decay in an
environment called a cloud chamber. Video 3.1.5 is an opportunity to learn about cloud chambers and to view an interesting Cloud
Chamber Demonstration from the Jefferson Lab.
3.1.7 https://chem.libretexts.org/@go/page/217253
How to Build a Cloud Chamber!
Figure 3.1.5 : Alpha particles, which are attracted to the negative plate and deflected by a relatively small amount, must be
positively charged and relatively massive. Beta particles, which are attracted to the positive plate and deflected a relatively large
amount, must be negatively charged and relatively light. Gamma rays, which are unaffected by the electric field, must be
uncharged.
Alpha (α) decay is the emission of an α particle from the nucleus. For example, polonium-210 undergoes α decay:
210 4 206 210 4 206
Po ⟶ 2
He + 82
Pb or Po ⟶ α+ 82
Pb (3.1.16)
84 84 2
Alpha decay occurs primarily in heavy nuclei (A > 200, Z > 83). Because the loss of an α particle gives a daughter nuclide with a
mass number four units smaller and an atomic number two units smaller than those of the parent nuclide, the daughter nuclide has a
larger n:p ratio than the parent nuclide. If the parent nuclide undergoing α decay lies below the band of stability, the daughter
nuclide will lie closer to the band.
Beta (β) decay is the emission of an electron from a nucleus. Iodine-131 is an example of a nuclide that undergoes β decay:
131 0 131 131 0 131
53
I ⟶ e+ 54
X or 53
I ⟶ −1
β+ 54
Xe (3.1.17)
−1
Beta decay, which can be thought of as the conversion of a neutron into a proton and a β particle, is observed in nuclides with a
large n:p ratio. The beta particle (electron) emitted is from the atomic nucleus and is not one of the electrons surrounding the
nucleus. Such nuclei lie above the band of stability. Emission of an electron does not change the mass number of the nuclide but
does increase the number of its protons and decrease the number of its neutrons. Consequently, the n:p ratio is decreased, and the
daughter nuclide lies closer to the band of stability than did the parent nuclide.
3.1.8 https://chem.libretexts.org/@go/page/217253
Gamma emission (γ emission) is observed when a nuclide is formed in an excited state and then decays to its ground state with the
emission of a γ ray, a quantum of high-energy electromagnetic radiation. The presence of a nucleus in an excited state is often
indicated by an asterisk (*). Cobalt-60 emits γ radiation and is used in many applications including cancer treatment:
60 ∗ 0 60
Co ⟶ γ+ Co (3.1.18)
27 0 27
There is no change in mass number or atomic number during the emission of a γ ray unless the γ emission accompanies one of the
other modes of decay.
Positron emission (β+ decay) is the emission of a positron from the nucleus. Oxygen-15 is an example of a nuclide that undergoes
positron emission:
15 0 15 15 0 15
8
O ⟶ e+ N or 8
O ⟶ +1
β+ N (3.1.19)
+1 7 7
Positron emission is observed for nuclides in which the n:p ratio is low. These nuclides lie below the band of stability. Positron
decay is the conversion of a proton into a neutron with the emission of a positron. The n:p ratio increases, and the daughter nuclide
lies closer to the band of stability than did the parent nuclide.
Electron capture occurs when one of the inner electrons in an atom is captured by the atom’s nucleus. For example, potassium-40
undergoes electron capture:
40 0 40
19
K+ e ⟶ 18
Ar (3.1.20)
−1
Electron capture occurs when an inner shell electron combines with a proton and is converted into a neutron. The loss of an inner
shell electron leaves a vacancy that will be filled by one of the outer electrons. As the outer electron drops into the vacancy, it will
emit energy. In most cases, the energy emitted will be in the form of an X-ray. Like positron emission, electron capture occurs for
“proton-rich” nuclei that lie below the band of stability. Electron capture has the same effect on the nucleus as does positron
emission: The atomic number is decreased by one and the mass number does not change. This increases the n:p ratio, and the
daughter nuclide lies closer to the band of stability than did the parent nuclide. Whether electron capture or positron emission
occurs is difficult to predict. The choice is primarily due to kinetic factors, with the one requiring the smaller activation energy
being the one more likely to occur. Figure 3.1.3 summarizes these types of decay, along with their equations and changes in atomic
and mass numbers.
Figure 3.1.6 : This table summarizes the type, nuclear equation, representation, and any changes in the mass or atomic numbers
for various types of decay.
PET Scan
Positron emission tomography (PET) scans use radiation to diagnose and track health conditions and monitor medical
treatments by revealing how parts of a patient’s body function (Figure 3.1.7). To perform a PET scan, a positron-emitting
3.1.9 https://chem.libretexts.org/@go/page/217253
radioisotope is produced in a cyclotron and then attached to a substance that is used by the part of the body being investigated.
This “tagged” compound, or radiotracer, is then put into the patient (injected via IV or breathed in as a gas), and how it is used
by the tissue reveals how that organ or other area of the body functions.
Figure 3.1.7 : A PET scanner (a) uses radiation to provide an image of how part of a patient’s body functions. The scans it
produces can be used to image a healthy brain (b) or can be used for diagnosing medical conditions such as Alzheimer’s disease
(c). (credit a: modification of work by Jens Maus)</
For example, F-18 is produced by proton bombardment of 18O ( O + p ⟶ F + n) and incorporated into a glucose
18
8
1
1
18
9
1
0
analog called fludeoxyglucose (FDG). How FDG is used by the body provides critical diagnostic information; for example,
since cancers use glucose differently than normal tissues, FDG can reveal cancers. The 18F emits positrons that interact with
nearby electrons, producing a burst of gamma radiation. This energy is detected by the scanner and converted into a detailed,
three-dimensional, color image that shows how that part of the patient’s body functions. Different levels of gamma radiation
produce different amounts of brightness and colors in the image, which can then be interpreted by a radiologist to reveal what is
going on. PET scans can detect heart damage and heart disease, help diagnose Alzheimer’s disease, indicate the part of a brain
that is affected by epilepsy, reveal cancer, show what stage it is, and how much it has spread, and whether treatments are
effective. Unlike magnetic resonance imaging and X-rays, which only show how something looks, the big advantage of PET
scans is that they show how something functions. PET scans are now usually performed in conjunction with a computed
tomography scan.
3.1.10 https://chem.libretexts.org/@go/page/217253
Figure 3.1.8 : Uranium-238 undergoes a radioactive decay series consisting of 14 separate steps before producing stable lead-
206. This series consists of eight α decays and six β decays.
Radioactive Half-Lives
Radioactive decay follows first-order kinetics. Since first-order reactions have already been covered in detail in the kinetics
chapter, we will now apply those concepts to nuclear decay reactions. Each radioactive nuclide has a characteristic, constant half-
life (t1/2), the time required for half of the atoms in a sample to decay. An isotope’s half-life allows us to determine how long a
sample of a useful isotope will be available, and how long a sample of an undesirable or dangerous isotope must be stored before it
decays to a low-enough radiation level that is no longer a problem.
For example, cobalt-60, an isotope that emits gamma rays used to treat cancer, has a half-life of 5.27 years (Figure 3.1.6). In a
given cobalt-60 source, since half of the Co nuclei decay every 5.27 years, both the amount of material and the intensity of the
60
27
radiation emitted is cut in half every 5.27 years. (Note that for a given substance, the intensity of radiation that it produces is
directly proportional to the rate of decay of the substance and the amount of the substance.) This is as expected for a process
following first-order kinetics. Thus, a cobalt-60 source that is used for cancer treatment must be replaced regularly to continue to be
effective.
Figure 3.1.9 : For cobalt-60, which has a half-life of 5.27 years, 50% remains after 5.27 years (one half-life), 25% remains after
10.54 years (two half-lives), 12.5% remains after 15.81 years (three half-lives), and so on.
Since nuclear decay follows first-order kinetics, we can adapt the mathematical relationships used for first-order chemical
reactions. We generally substitute the number of nuclei, N, for the concentration. If the rate is stated in nuclear decays per second,
we refer to it as the activity of the radioactive sample. The rate for radioactive decay is:
3.1.11 https://chem.libretexts.org/@go/page/217253
decay rate = λN (3.1.21)
where N0 is the initial number of nuclei or moles of the isotope, and Nt is the number of nuclei/moles remaining at time t. We will
not concern ourselves with the calculation of half-life in this course.
Because each nuclide has a specific number of nucleons, a particular balance of repulsion and attraction, and its own degree of
stability, the half-lives of radioactive nuclides vary widely. For example: the half-life of Bi is 1.9 × 1019 years; Ra is 24,000
209
83
239
94
years; Rn is 3.82 days; and element-111 (Rg for roentgenium) is 1.5 × 10–3 seconds. The half-lives of a number of radioactive
222
86
isotopes important to medicine are shown in Table 3.1.1, and others are listed in Appendix N1.
Table 3.1.2: Half-lives of Radioactive Isotopes Important to Medicine
Type Decay Mode Half-Life Uses
Radiometric Dating
Several radioisotopes have half-lives and other properties that make them useful for purposes of “dating” the origin of objects such
as archaeological artifacts, formerly living organisms, or geological formations. This process is radiometric dating and has been
responsible for many breakthrough scientific discoveries about the geological history of the earth, the evolution of life, and the
history of human civilization. We will explore some of the most common types of radioactive dating and how the particular
isotopes work for each type.
Radioactive Dating Using Carbon-14
The radioactivity of carbon-14 provides a method for dating objects that were a part of a living organism. This method of
radiometric dating, which is also called radiocarbon dating or carbon-14 dating, is accurate for dating carbon-containing substances
that are up to about 30,000 years old, and can provide reasonably accurate dates up to a maximum of about 50,000 years old.
Naturally occurring carbon consists of three isotopes: C, which constitutes about 99% of the carbon on earth; C, about 1% of
12
6
13
6
the total; and trace amounts of C. Carbon-14 forms in the upper atmosphere by the reaction of nitrogen atoms with neutrons from
14
6
14 12
All isotopes of carbon react with oxygen to produce CO2 molecules. The ratio of CO to CO depends on the ratio of CO to
6 2 6 2
14
6
CO in the atmosphere. The natural abundance of CO in the atmosphere is approximately 1 part per trillion; until recently, this
12 14
6 6
14 12
has generally been constant over time, as seen is gas samples found trapped in ice. The incorporation of C CO and CO 14
6 6 2 6 2
into plants is a regular part of the photosynthesis process, which means that the C : C ratio found in a living plant is the same
14
6
12
3.1.12 https://chem.libretexts.org/@go/page/217253
as the C : C ratio in the atmosphere. But when the plant dies, it no longer traps carbon through photosynthesis. Because C is
14
6
12
6
12
6
a stable isotope and does not undergo radioactive decay, its concentration in the plant does not change. However, carbon-14 decays
by β emission with a half-life of 5730 years:
14 14 0
6
C ⟶ 7
N+ e (3.1.25)
−1
Thus, the C : C ratio gradually decreases after the plant dies. The decrease in the ratio with time provides a measure of the time
14
6
12
6
that has elapsed since the death of the plant (or other organism that ate the plant). Figure 3.1.10 visually depicts this process.
Figure 3.1.10 : Along with stable carbon-12, radioactive carbon-14 is taken in by plants and animals, and remains at a constant
level within them while they are alive. After death, the C-14 decays and the C-14:C-12 ratio in the remains decreases. Comparing
this ratio to the C-14:C-12 ratio in living organisms allows us to determine how long ago the organism lived (and died).
For example, with the half-life of C being 5730 years, if the C : C ratio in a wooden object found in an archaeological dig is
14
6
14
6
12
6
half what it is in a living tree, this indicates that the wooden object is 5730 years old. Highly accurate determinations of C : C14
6
12
6
ratios can be obtained from very small samples (as little as a milligram) by the use of a mass spectrometer.
There have been some significant, well-documented changes to the C : C ratio. The accuracy of a straightforward application
14
6
12
6
of this technique depends on the C : C ratio in a living plant being the same now as it was in an earlier era, but this is not
14
6
12
6
12
always valid. Due to the increasing accumulation of CO2 molecules (largely CO ) in the atmosphere caused by combustion of
6 2
fossil fuels (in which essentially all of the C has decayed), the ratio of C : C in the atmosphere may be changing. This
14
6
14
6
12
6
12
manmade increase in CO in the atmosphere causes the C : C ratio to decrease, and this in turn affects the ratio in currently
6 2
14
6
12
6
living organisms on the earth. Fortunately, however, we can use other data, such as tree dating via examination of annual growth
rings, to calculate correction factors. With these correction factors, accurate dates can be determined. In general, radioactive dating
only works for about 10 half-lives; therefore, the limit for carbon-14 dating is about 57,000 years.
Radioactive Dating Using Nuclides Other than Carbon-14
Radioactive dating can also use other radioactive nuclides with longer half-lives to date older events. For example, uranium-238
(which decays in a series of steps into lead-206) can be used for establishing the age of rocks (and the approximate age of the oldest
rocks on earth). Since U-238 has a half-life of 4.5 billion years, it takes that amount of time for half of the original U-238 to decay
into Pb-206. In a sample of rock that does not contain appreciable amounts of Pb-208, the most abundant isotope of lead, we can
assume that lead was not present when the rock was formed. Therefore, by measuring and analyzing the ratio of U-238:Pb-206, we
can determine the age of the rock. This assumes that all of the lead-206 present came from the decay of uranium-238. If there is
additional lead-206 present, which is indicated by the presence of other lead isotopes in the sample, it is necessary to make an
adjustment. Potassium-argon dating uses a similar method. K-40 decays by positron emission and electron capture to form Ar-40
with a half-life of 1.25 billion years. If a rock sample is crushed and the amount of Ar-40 gas that escapes is measured,
determination of the Ar-40:K-40 ratio yields the age of the rock. Other methods, such as rubidium-strontium dating (Rb-87 decays
3.1.13 https://chem.libretexts.org/@go/page/217253
into Sr-87 with a half-life of 48.8 billion years), operate on the same principle. To estimate the lower limit for the earth’s age,
scientists determine the age of various rocks and minerals, making the assumption that the earth is older than the oldest rocks and
minerals in its crust. As of 2014, the oldest known rocks on earth are the Jack Hills zircons from Australia, found by uranium-lead
dating to be almost 4.4 billion years old.
Summary
Video 3.1.6 : A video summary of some of the nuclear chemistry you just learned about.
An atomic nucleus consists of protons and neutrons, collectively called nucleons. Although protons repel each other, the nucleus is
held tightly together by a short-range, but very strong, force called the strong nuclear force. A nucleus has less mass than the total
mass of its constituent nucleons. This “missing” mass is the mass defect, which has been converted into the binding energy that
holds the nucleus together according to Einstein’s mass-energy equivalence equation, E = mc2. Of the many nuclides that exist,
only a small number are stable. Nuclides with even numbers of protons or neutrons, or those with magic numbers of nucleons, are
especially likely to be stable. These stable nuclides occupy a narrow band of stability on a graph of number of protons versus
number of neutrons. The binding energy per nucleon is largest for the elements with mass numbers near 56; these are the most
stable nuclei.
Nuclei can undergo reactions that change their number of protons, number of neutrons, or energy state. Many different particles can
be involved in nuclear reactions. The most common are protons, neutrons, positrons (which are positively charged electrons), alpha
(α) particles (which are high-energy helium nuclei), beta (β) particles (which are high-energy electrons), and gamma (γ) rays
(which compose high-energy electromagnetic radiation). As with chemical reactions, nuclear reactions are always balanced. When
a nuclear reaction occurs, the total mass (number) and the total charge remain unchanged.
Nuclei that have unstable n:p ratios undergo spontaneous radioactive decay. The most common types of radioactivity are α decay, β
decay, γ emission, positron emission, and electron capture. Nuclear reactions also often involve γ rays, and some nuclei decay by
electron capture. Each of these modes of decay leads to the formation of a new nucleus with a more stable n:p ratio. Some
substances undergo radioactive decay series, proceeding through multiple decays before ending in a stable isotope. All nuclear
decay processes follow first-order kinetics, and each radioisotope has its own characteristic half-life, the time that is required for
half of its atoms to decay. Because of the large differences in stability among nuclides, there is a very wide range of half-lives of
radioactive substances. Many of these substances have found useful applications in medical diagnosis and treatment, determining
the age of archaeological and geological objects, and more.
Key Equations
E = mc2
decay rate = λN
ln 2 0.693
t1/2 = =
λ λ
Glossary
3.1.14 https://chem.libretexts.org/@go/page/217253
alpha particle
(α or He or α) high-energy helium nucleus; a helium atom that has lost two electrons and contains two protons and two
4
2
4
2
neutrons
antimatter
particles with the same mass but opposite properties (such as charge) of ordinary particles
band of stability
(also, belt of stability, zone of stability, or valley of stability) region of graph of number of protons versus number of neutrons
containing stable (nonradioactive) nuclides
beta (β) decay
breakdown of a neutron into a proton, which remains in the nucleus, and an electron, which is emitted as a beta particle
beta particle
(β or −1
0
e or −1
0
) high-energy electron
β
half-life (t1/2)
time required for half of the atoms in a radioactive sample to decay
magic number
nuclei with specific numbers of nucleons that are within the band of stability
mass defect
difference between the mass of an atom and the summed mass of its constituent subatomic particles (or the mass “lost” when
nucleons are brought together to form a nucleus)
mass-energy equivalence equation
Albert Einstein’s relationship showing that mass and energy are equivalent
nuclear binding energy
energy lost when an atom’s nucleons are bound together (or the energy needed to break a nucleus into its constituent protons
and neutrons)
nuclear chemistry
study of the structure of atomic nuclei and processes that change nuclear structure
nuclear reaction
change to a nucleus resulting in changes in the atomic number, mass number, or energy state
nucleon
collective term for protons and neutrons in a nucleus
nuclide
nucleus of a particular isotope
parent nuclide
unstable nuclide that changes spontaneously into another (daughter) nuclide
3.1.15 https://chem.libretexts.org/@go/page/217253
positron ( +1
0
β or 0
+1
)
e
antiparticle to the electron; it has identical properties to an electron, except for having the opposite (positive) charge
positron emission
(also, β+ decay) conversion of a proton into a neutron, which remains in the nucleus, and a positron, which is emitted
radioactive decay
spontaneous decay of an unstable nuclide into another nuclide
radioactivity
phenomenon exhibited by an unstable nucleon that spontaneously undergoes change into a nucleon that is more stable; an
unstable nucleon is said to be radioactive
radiocarbon dating
highly accurate means of dating objects 30,000–50,000 years old that were derived from once-living matter; achieved by
calculating the ratio of C : C in the object vs. the ratio of C : C in the present-day atmosphere
14
6
12
6
14
6
12
6
radioisotope
isotope that is unstable and undergoes conversion into a different, more stable isotope
radiometric dating
use of radioisotopes and their properties to date the formation of objects such as archeological artifacts, formerly living
organisms, or geological formations
strong nuclear force
force of attraction between nucleons that holds a nucleus together
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
TED-Ed’s commitment to creating lessons worth sharing is an extension of TED’s mission of spreading great ideas. Within
TED-Ed’s growing library of TED-Ed animations, you will find carefully curated educational videos, many of which represent
collaborations between talented educators and animators nominated through the TED-Ed website.
Fuse School, Open Educational Resource free of charge, under a Creative Commons License: Attribution-NonCommercial CC
BY-NC (View License Deed: https://creativecommons.org/licenses/by-nc/4.0/)
3.1: Nuclear Chemistry and Radioactive Decay is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
3.1.16 https://chem.libretexts.org/@go/page/217253
3.1.1: Practice Problems- Nuclear Chemistry and Radioactive Decay (Optional)
PROBLEM 3.1.1.1
Write the following isotopes in nuclide notation (e.g., " 14
6
C ")
a. oxygen-14
b. copper-70
c. tantalum-175
d. francium-217
Answer a
14
O
8
Answer b
70
29
Cu
Answer c
175
Ta
73
Answer d
217
Fr
87
Problem 3.1.1
PROBLEM 3.1.1.2
For the following isotopes that have missing information, fill in the missing information to complete the notation
a. 34
14
X
b. 36
P
c. 57
X
Mn
d. 121
56
X
Answer a
34
14
Si
Answer b
36
15
P
3.1.1.1 https://chem.libretexts.org/@go/page/217256
Answer c
57
Mn
25
Answer d
121
56
Ba
PROBLEM 3.1.1.3
Write the nuclide notation, including charge if applicable, for atoms with the following characteristics:
a. 25 protons, 20 neutrons, 24 electrons
b. 45 protons, 24 neutrons, 43 electrons
c. 53 protons, 89 neutrons, 54 electrons
d. 97 protons, 146 neutrons, 97 electrons
Answer a
45 +1
Mn
25
Answer b
69 +2
Rh
45
Answer c
142 −1
I
53
Answer d
243
Bk
97
Problem 3.1.3
PROBLEM 3.1.1.4
Which of the following nuclei lie within the band of stability?
a. chlorine-37
b. calcium-40
c. 204Bi
d. 56Fe
e. 206Pb
f. 211Pb
g. 222Rn
3.1.1.2 https://chem.libretexts.org/@go/page/217256
h. carbon-14
Answer
(a), (b), (c), (d), and (e)
PROBLEM 3.1.1.5
Which of the following nuclei lie within the band of stability?
a. argon-40
b. oxygen-16
c. 122Ba
d. 58Ni
e. 205Tl
f. 210Tl
g. 226Ra
h. magnesium-24
Answer
(b), (e - very close), and (h)
Problem 3.1.5
PROBLEM 3.1.1.6
Write a brief description or definition of each of the following:
a. nucleon
b. α particle
c. β particle
d. positron
e. γ ray
f. nuclide
g. mass number
h. atomic number
Answer a
collective term for protons and neutrons in a nucleus
Answer b
3.1.1.3 https://chem.libretexts.org/@go/page/217256
(α or He or
4
2
4
2
α ) high-energy helium nucleus; a helium atom that has lost two electrons and contains two protons and two
neutrons
Answer c
( β or −1
0
e or −1
0
β) high-energy electron
Answer d
antiparticle to the electron; it has identical properties to an electron, except for having the opposite (positive) charge
Answer e
(γ or 0
0
γ ) short wavelength, high-energy electromagnetic radiation that exhibits wave-particle duality
Answer f
nucleus of a particular isotope
Answer g
sum of the numbers of neutrons and protons in the nucleus of an atom
Answer h
number of protons in the nucleus of an atom
PROBLEM 3.1.1.7
Complete each of the following equations by adding the missing species:
a. 27
13
4
Al + 2 He ⟶ ? +
1
0
n
b. 239
94
Pu+ ? ⟶
242
96
Cm +
1
0
n
c. 14
7
4
N + 2 He ⟶ ? + 1 H
1
d. 235
92
U ⟶ ?+
135
55
Cs + 4
1
0
n
Answer a
27 4 30 1
13
Al + 2 He ⟶ 15
P+ n
0
Answer b
239 4 242 1
Pu + 2 He ⟶ 96
Cm + n
94 0
Answer c
14 4 17 1
N+ He ⟶ O+ H
7 2 8 1
Answer d
235 96 135 1
92
U ⟶ 37
Rb + 55
Cs + 4 n
0
3.1.1.4 https://chem.libretexts.org/@go/page/217256
Problem 3.1.7
PROBLEM 3.1.1.8
Complete each of the following equations:
a. 7
3
Li + ? ⟶ 2
4
2
He
b. 14
6
C ⟶
14
7
N+ ?
c. 27
13
4
Al + 2 He ⟶ ? +
1
0
n
d. 250
96
Cm ⟶ ? + 38 Sr + 4
98 1
0
n
Answer a
7 1 4
Li + H ⟶ 2 He
3 1 2
Answer b
14 14 0
6
C ⟶ 7
N+ e
−1
Answer c
27 4 30 1
13
Al + 2 He ⟶ P+ n
15 0
Answer d
250 148 98 1
96
Cm ⟶ Ce + 38 Sr + 4 n
58 0
PROBLEM 3.1.1.9
Write a balanced equation for each of the following nuclear reactions:
a. the production of 17O from 14N by α particle bombardment
b. the production of 14C from 14N by neutron bombardment
c. the production of 233Th from 232Th by neutron bombardment
d. the production of 239U from 238U by H bombardment 2
1
Answer a
14 2 17 1
7
N + He ⟶ 8
O + 1H
Answer b
14 1 14 1
7
N+ n ⟶ 6
N + 1H
0
Answer c
3.1.1.5 https://chem.libretexts.org/@go/page/217256
232 1 233
90
Th + n ⟶ 90
Th
0
Answer d
238 2 239 1
92
U + 1H ⟶ 92
U + 1H
Problem 3.1.9
PROBLEM 3.1.1.10
Technetium-99 is prepared from 98Mo. Molybdenum-98 combines with a neutron to give molybdenum-99, an unstable isotope
that emits a β particle to yield an excited form of technetium-99, represented as 99Tc*. This excited nucleus relaxes to the ground
state, represented as 99Tc, by emitting a γ ray. The ground state of 99Tc then emits a β particle. Write the equations for each of
these nuclear reactions.
Answer
98 1 99
42
Mo + n → 42
Mo
0
99 0 99 ∗
42
Mo → e+ Tc
−1 43
∗
99 0 99
Tc → γ+ Tc
43 0 43
99 0 99
Tc → e+ Ru
43 −1 44
PROBLEM 3.1.1.11
What changes occur to the atomic number and mass of a nucleus during each of the following decay scenarios?
a. an α particle is emitted
b. a β particle is emitted
c. γ radiation is emitted
d. a positron is emitted
e. an electron is captured
Answer a
Since an α particle is the same as a 4
2
He nucleus, the mass number will decrease by 4 and the atomic number will decrease by
2.
Answer b
Since a β particle is the same as 0
−1
, the mass number will not change but the atomic number will increase by 1.
e
3.1.1.6 https://chem.libretexts.org/@go/page/217256
Answer c
Since a γ ray has no mass (it is energy) the mass number and atomic number do not change.
Answer d
A positron is the opposite of a β particle, it is 0
+1
, the mass number will not change but the atomic number will decrease by
e
Answer e
Electron capture has the same effect on the nucleus as positron emission: The atomic number is decreased by one and the
mass number does not change.
PROBLEM 3.1.1.12
What is the change in the nucleus that results from the following decay scenarios?
a. emission of a β particle
b. emission of a β+ particle
c. capture of an electron
Answer a
conversion of a neutron to a proton: 1
0
n ⟶
1
1
p+
+1
0
e
Answer b
conversion of a proton to a neutron; the positron has the same mass as an electron and the same magnitude of positive charge
as the electron has negative charge
when the n:p ratio of a nucleus is too low, a proton is converted into a neutron with the emission of a positron:
1 1 0
p ⟶ n+ e
1 0 +1
Answer c
In a proton-rich nucleus, an inner atomic electron can be absorbed. In simplest form, this changes a proton into a neutron:
1 0 1
p+ e ⟶ p
1 −1 0
PROBLEM 3.1.1.13
Explain how unstable heavy nuclides (atomic number > 83) may decompose to form
nuclides of greater stability if
(a) they are below the band of stability and
(b) they are above the band of stability
Answer
Nuclei below the band of stability will undergo positron decay, while those above the band of stability will undergo beta
decay. Heavy nuclei past the band of stability will undergo alpha decay
3.1.1.7 https://chem.libretexts.org/@go/page/217256
PROBLEM 3.1.1.14
Which of the following nuclei is most likely to decay by positron emission? Explain your choice.
a. chromium-53
b. manganese-51
c. iron-59
Answer
29 26
Manganese-51 is most likely to decay by positron emission. The n:p ratio for Cr-53 is = 1.21; for Mn-51, it is =
24 25
33
1.04; for Fe-59, it is = 1.27. Positron decay occurs when the n:p ratio is low. Mn-51 has the lowest n:p ratio and
26
PROBLEM 3.1.1.15
The following nuclei do not lie in the band of stability. How would they be expected to
decay? Explain your answer.
a. 34
15
P
b. 239
92
U
c. 38
20
Ca
d. 3
1
H
e. 245
94
Pu
Answer a
Above the band of stability, beta decay is expected
Answer b
Beyond the band of stability, heavy nuclei undergo alpha decay
Answer c
Below the band of stability, positron decay is expected
Answer d
Above the band of stability, beta decay is expected
Answer e
Beyond the band of stability, heavy nuclei undergo alpha decay
PROBLEM 3.1.1.16
Write a nuclear reaction for each step in the formation of Po from
218
84
U , which proceeds by a series of decay reactions
238
92
Answer
238 234 4
U ⟶ 90
Th + 2 He
92
234 234 0
Th ⟶ Pa + e
90 91 −1
3.1.1.8 https://chem.libretexts.org/@go/page/217256
234 234 0
91
Pa ⟶ U+ e
92 −1
234 230 4
92
U ⟶ 90
Th + 2 He
230 226 4
90
Th ⟶ 88
Ra + 2 He
226 222 4
88
Ra ⟶ Rn + 2 He
86
222 218 4
86
Rn ⟶ Po + 2 He
84
PROBLEM 3.1.1.17
Write a nuclear reaction for each step in the formation of Pb from
208
82
228
90
Th, which proceeds by a series of decay reactions
Answer
228 4 224
90
Th → 2
He + 88
Ra
224 4 220
Ra → He + Rn
88 2 86
220 4 216
Th → He + Po
86 2 84
216 4 212
84
Po → 2
He + 82
Pb
212 0 212
82
Pb → e+ 83
Bi
−1
212 0 212
83
Bi → e+ 84
Po
−1
212 4 208
Po → He + Pb
84 2 82
Problem 3.1.17
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
3.1.1: Practice Problems- Nuclear Chemistry and Radioactive Decay (Optional) is shared under a not declared license and was authored, remixed,
and/or curated by LibreTexts.
3.1.1.9 https://chem.libretexts.org/@go/page/217256
3.2: Organization of the Periodic Table
Skills to Develop
State the periodic law and explain the organization of elements in the periodic table
Predict the general properties of elements based on their location within the periodic table
Identify metals, nonmetals, and metalloids by their properties and/or location on the periodic table
3.2.1 https://chem.libretexts.org/@go/page/217254
Figure 3.2.1 : (a) Dimitri Mendeleev is widely credited with creating (b) the first periodic table of the elements. (credit a:
modification of work by Serge Lachinov; credit b: modification of work by “Den fjättrade ankan”/Wikimedia Commons)
Video 3.2.2 : An introduction to the organization of the periodic table into periods and groups. This organization will be important
as we continue building on the principles of chemistry.
By the twentieth century, it became apparent that the periodic relationship involved atomic numbers rather than atomic masses. The
modern statement of this relationship, the periodic law, is as follows: the properties of the elements are periodic functions of their
atomic numbers. A modern periodic table arranges the elements in increasing order of their atomic numbers and groups atoms with
similar properties in the same vertical column (Figure 3.2.2). Each box represents an element and contains its atomic number,
symbol, average atomic mass, and (sometimes) name. The elements are arranged in seven horizontal rows, called periods or series,
and 18 vertical columns, called groups. Groups are labeled at the top of each column. In the United States, the labels traditionally
were numerals with capital letters. However, IUPAC recommends that the numbers 1 through 18 be used, and these labels are more
common. For the table to fit on a single page, parts of two of the rows, a total of 14 columns, are usually written below the main
body of the table.
Figure 3.2.2 : Elements in the periodic table are organized according to their properties.
3.2.2 https://chem.libretexts.org/@go/page/217254
Many elements differ dramatically in their chemical and physical properties, but some elements are similar in their behaviors. For
example, many elements appear shiny, are malleable (able to be deformed without breaking) and ductile (can be drawn into wires),
and conduct heat and electricity well. Other elements are not shiny, malleable, or ductile, and are poor conductors of heat and
electricity. We can sort the elements into large classes with common properties: metals (elements that are shiny, malleable, good
conductors of heat and electricity—shaded yellow); nonmetals (elements that appear dull, poor conductors of heat and electricity—
shaded green); and metalloids (elements that conduct heat and electricity moderately well, and possess some properties of metals
and some properties of nonmetals—shaded purple).
The elements can also be classified into the main-group elements (or representative elements) in the columns labeled 1, 2, and 13–
18; the transition metals in the columns labeled 3–12; and inner transition metals in the two rows at the bottom of the table (the top-
row elements are called lanthanides and the bottom-row elements are actinides; Figure 3.2.3). The elements can be subdivided
further by more specific properties, such as the composition of the compounds they form. For example, the elements in group 1 (the
first column) form compounds that consist of one atom of the element and one atom of hydrogen. These elements (except
hydrogen) are known as alkali metals, and they all have similar chemical properties. The elements in group 2 (the second column)
form compounds consisting of one atom of the element and two atoms of hydrogen: These are called alkaline earth metals, with
similar properties among members of that group. Other groups with specific names are the pnictogens (group 15), chalcogens
(group 16), halogens (group 17), and the noble gases (group 18, also known as inert gases). The groups can also be referred to by
the first element of the group: For example, the chalcogens can be called the oxygen group or oxygen family. Hydrogen is a unique,
nonmetallic element with properties similar to both group 1 and group 17 elements. For that reason, hydrogen may be shown at the
top of both groups, or by itself.
Figure 3.2.3 : The periodic table organizes elements with similar properties into groups.
3.2.3 https://chem.libretexts.org/@go/page/217254
a. halogen
b. alkaline earth metal
c. alkali metal
d. chalcogen
Exercise 3.2.1
Give the group name for each of the following elements:
a. krypton
b. selenium
c. barium
d. lithium
Answer a
noble gas
Answer b
chalcogen
Answer c
alkaline earth metal
Answer d
alkali metal
In studying the periodic table, you might have noticed something about the atomic masses of some of the elements. Element 43
(technetium), element 61 (promethium), and most of the elements with atomic number 84 (polonium) and higher have their atomic
mass given in square brackets. This is done for elements that consist entirely of unstable, radioactive isotopes (you will learn more
about radioactivity in the nuclear chemistry chapter). An average atomic weight cannot be determined for these elements because
their radioisotopes may vary significantly in relative abundance, depending on the source, or may not even exist in nature. The
number in square brackets is the atomic mass number (and approximate atomic mass) of the most stable isotope of that element.
Summary
Video 3.2.3 : A summary of the discovery and properties of the periodic table.
The discovery of the periodic recurrence of similar properties among the elements led to the formulation of the periodic table, in
which the elements are arranged in order of increasing atomic number in rows known as periods and columns known as groups.
Elements in the same group of the periodic table have similar chemical properties. Elements can be classified as metals, metalloids,
and nonmetals, or as a main-group elements, transition metals, and inner transition metals. Groups are numbered 1–18 from left to
right. The elements in group 1 are known as the alkali metals; those in group 2 are the alkaline earth metals; those in 15 are the
pnictogens; those in 16 are the chalcogens; those in 17 are the halogens; and those in 18 are the noble gases.
3.2.4 https://chem.libretexts.org/@go/page/217254
Glossary
actinide
inner transition metal in the bottom of the bottom two rows of the periodic table
alkali metal
element in group 1
chalcogen
element in group 16
group
vertical column of the periodic table
halogen
element in group 17
inert gas
(also, noble gas) element in group 18
lanthanide
inner transition metal in the top of the bottom two rows of the periodic table
main-group element
(also, representative element) element in columns 1, 2, and 12–18
metal
element that is shiny, malleable, good conductor of heat and electricity
metalloid
element that conducts heat and electricity moderately well, and possesses some properties of metals and some properties of
nonmetals
noble gas
(also, inert gas) element in group 18
nonmetal
element that appears dull, poor conductor of heat and electricity
period
(also, series) horizontal row of the periodic table
periodic law
properties of the elements are periodic function of their atomic numbers.
periodic table
table of the elements that places elements with similar chemical properties close together
pnictogen
3.2.5 https://chem.libretexts.org/@go/page/217254
element in group 15
representative element
(also, main-group element) element in columns 1, 2, and 12–18
transition metal
element in columns 3–11
series
(also, period) horizontal row of the period table
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
TED-Ed’s commitment to creating lessons worth sharing is an extension of TED’s mission of spreading great ideas. Within
TED-Ed’s growing library of TED-Ed animations, you will find carefully curated educational videos, many of which represent
collaborations between talented educators and animators nominated through the TED-Ed website.
Fuse School, Open Educational Resource free of charge, under a Creative Commons License: Attribution-NonCommercial CC
BY-NC (View License Deed: https://creativecommons.org/licenses/by-nc/4.0/)
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
3.2: Organization of the Periodic Table is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
3.2.6 https://chem.libretexts.org/@go/page/217254
3.2.1: Practice Problems - Organization of the Periodic Table (Optional)
PROBLEM 3.2.1.1
Using the periodic table, classify each of the following elements as a metal or a nonmetal, and then further classify each as a
main-group element, transition metal, or inner transition metal:
a. uranium
b. bromine
c. strontium
d. neon
e. gold
f. americium
g. rhodium
h. sulfur
i. carbon
j. potassium
Answer a
metal; inner transition metal
Answer b
nonmetal, main group element
Answer c
metal, main group element
Answer d
nonmetal, main group element
Answer e
metal, transition metal
Answer f
metal, inner transition metal
Answer g
metal, transition metal
Answer h
nonmetal, main group element
Answer i
nonmetal, main group element
Answer j
metal, main group element
PROBLEM 3.2.1.2
Using the periodic table, classify each of the following elements as a metal or a nonmetal, and then further classify each as a
main-group (representative) element, transition metal, or inner transition metal:
3.2.1.1 https://chem.libretexts.org/@go/page/217257
a. cobalt
b. europium
c. iodine
d. indium
e. lithium
f. oxygen
g. cadmium
h. terbium
i. rhenium
Answer a
metal, transition metal
Answer b
metal, inner transition metal
Answer c
nonmetal, main group element
Answer d
metal, main group element
Answer e
metal, main group element
Answer f
nonmetal, main group element
Answer g
metal, transition metal
Answer h
metal, inner transition metal
Answer i
metal, transition metal
PROBLEM 3.2.1.3
Using the periodic table, identify the lightest member of each of the following groups:
a. noble gases
b. alkaline earth metals
c. alkali metals
d. chalcogens
Answer a
He
Answer b
Be
Answer c
3.2.1.2 https://chem.libretexts.org/@go/page/217257
Li
Answer d
O
PROBLEM 3.2.1.4
Using the periodic table, identify the heaviest member of each of the following groups:
a. alkali metals
b. chalcogens
c. noble gases
d. alkaline earth metals
Answer a
Fr
Answer b
Po
Answer c
Rn
Answer d
Ra
PROBLEM 3.2.1.5
Use the periodic table to give the name and symbol for each of the following elements:
(a) the noble gas in the same period as germanium
(b) the alkaline earth metal in the same period as selenium
(c) the halogen in the same period as lithium
(d) the chalcogen in the same period as cadmium
Answer a
Kr; Krypton
Answer b
Ca; Calcium
Answer c
F; Fluorine
Answer d
Te; Tellurium
PROBLEM 3.2.1.6
Use the periodic table to give the name and symbol for each of the following elements:
(a) the halogen in the same period as the alkali metal with 11 protons
(b) the alkaline earth metal in the same period with the neutral noble gas with 18 electrons
(c) the noble gas in the same row as an isotope with 30 neutrons and 25 protons
(d) the noble gas in the same period as gold
3.2.1.3 https://chem.libretexts.org/@go/page/217257
Answer a
Cl; Chlorine
Answer b
Mg; Magnesium
Answer c
Kr; Krypton
Answer d
Rn; Radon
PROBLEM 3.2.1.7
Write a symbol for each of the following neutral isotopes. Include the atomic number and mass number for each.
(a) the alkali metal with 11 protons and a mass number of 23
(b) the noble gas element with and 75 neutrons in its nucleus and 54 electrons in the neutral atom
(c) the isotope with 33 protons and 40 neutrons in its nucleus
(d) the alkaline earth metal with 88 electrons and 138 neutrons
Answer a
23
Na
11
Answer b
129
Xe
54
Answer c
73
As
33
Answer d
226
Ra
88
PROBLEM 3.2.1.8
Write a symbol for each of the following neutral isotopes. Include the atomic number and mass number for each.
a. the chalcogen with a mass number of 125
b. the halogen whose longest-lived isotope is radioactive
c. the noble gas, used in lighting, with 10 electrons and 10 neutrons
d. the lightest alkali metal with three neutrons
Answer a
125
52
Te
Answer b
210
85
At
Answer c
20
10
Ne
Answer d
6
Li
3
3.2.1.4 https://chem.libretexts.org/@go/page/217257
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
3.2.1: Practice Problems - Organization of the Periodic Table (Optional) is shared under a not declared license and was authored, remixed, and/or
curated by LibreTexts.
3.2.1.5 https://chem.libretexts.org/@go/page/217257
3.3: Predicting Charges of Ions
Skills to Develop
Define ions, cations and anions
Predict the charge of an anion or cation based upon the location of the element in the periodic table
Ions of Elements
In ordinary chemical reactions, the nucleus of each atom remains unchanged. Electrons, however, can be added to atoms by transfer
form other atoms, lost by transfer to other atoms, or shared with other atoms. The transfer and sharing of electrons among atoms
govern the chemistry of the elements. During the formation of some compounds, atoms gain or lose electrons, and form electrically
charged particles called ions (Figure 3.3.1; Video 3.3.1)
Figure 3.3.1 : (a) A sodium atom (Na) has equal numbers of protons and electrons (11) and is uncharged. (b) A sodium cation
(Na+) has lost an electron, so it has one more proton (11) than electrons (10), giving it an overall positive charge, signified by a
superscripted plus sign.
One can use the periodic table to predict whether an atom will form an anion or a cation, and you can often predict the charge of
the resulting ion. Atoms of many main-group metals lose enough electrons to leave them with the same number of electrons as an
atom of the preceding noble gas. To illustrate, an atom of an alkali metal (group 1) loses one electron and forms a cation with a 1+
charge; an alkaline earth metal (group 2) loses two electrons and forms a cation with a 2+ charge, and so on. For example, a neutral
calcium atom, with 20 protons and 20 electrons, readily loses two electrons. This results in a cation with 20 protons, 18 electrons,
and a 2+ charge. It has the same number of electrons as atoms of the preceding noble gas, argon, and is symbolized Ca2+. The name
of a metal ion is the same as the name of the metal atom from which it forms, so Ca2+ is called a calcium ion.
When atoms of nonmetal elements form ions, they generally gain enough electrons to give them the same number of electrons as an
atom of the next noble gas in the periodic table. Atoms of group 17 gain one electron and form anions with a 1− charge; atoms of
group 16 gain two electrons and form ions with a 2− charge, and so on. For example, the neutral bromine atom, with 35 protons
and 35 electrons, can gain one electron to provide it with 36 electrons. This results in an anion with 35 protons, 36 electrons, and a
1− charge. It has the same number of electrons as atoms of the next noble gas, krypton, and is symbolized Br−. (A discussion of the
theory supporting the favored status of noble gas electron numbers reflected in these predictive rules for ion formation is provided
in a later chapter of this text.)
Note the usefulness of the periodic table in predicting likely ion formation and charge (Figure 3.3.2). Moving from the far left to
the right on the periodic table, main-group elements tend to form cations with a charge equal to the group number. That is, group 1
elements form 1+ ions; group 2 elements form 2+ ions, and so on. Moving from the far right to the left on the periodic table,
elements often form anions with a negative charge equal to the number of groups moved left from the noble gases. For example,
group 17 elements (one group left of the noble gases) form 1− ions; group 16 elements (two groups left) form 2− ions, and so on.
This trend can be used as a guide in many cases, but its predictive value decreases when moving toward the center of the periodic
table. In fact, transition metals and some other metals often exhibit variable charges that are not predictable by their location in the
table. For example, copper can form ions with a 1+ or 2+ charge, and iron can form ions with a 2+ or 3+ charge.
3.3.1 https://chem.libretexts.org/@go/page/217255
Figure 3.3.2 : Some elements exhibit a regular pattern of ionic charge when they form ions.
Exercise 3.3.1
Give the symbol and name for the ion with 34 protons and 36 electrons.
Answer
Se2−, the selenide ion
Exercise 3.3.2
3.3.2 https://chem.libretexts.org/@go/page/217255
Aluminum and carbon react to form an ionic compound. Predict which forms an anion, which forms a cation, and the charges of
each ion. Write the symbol for each ion and name them.
Answer
Al will form a cation with a charge of 3+: Al3+, an aluminum ion. Carbon will form an anion with a charge of 4−: C4−, a
carbide ion.
Summary
Glossary
anion
negative ions that are formed when a nonmental atom gains one or more electrons.
cation
positive ions that are formed when an atom loses electrons
ion
atom or molecule that has gained or lost one or more of its valence electrons, giving it a net electrical charge.
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Fuse School, Open Educational Resource free of charge, under a Creative Commons License: Attribution-NonCommercial CC
BY-NC (View License Deed: https://creativecommons.org/licenses/by-nc/4.0/)
3.3.3 https://chem.libretexts.org/@go/page/217255
3.3: Predicting Charges of Ions is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
3.3.4 https://chem.libretexts.org/@go/page/217255
CHAPTER OVERVIEW
4: Molecules, Compounds and Quantifying Chemicals
4.1: Chemical Bonding
4.2: Representing ionic and molecular compounds and molecules
4.2.1: Practice Problems- Writing Chemical Formulas
4.3: Chemical Nomenclature
4.3.1: Practice Problems- Molecular and Ionic Compounds
4.4: Formula Mass, Percent Composition, and the Mole
4.4.1: Practice Problems- Formula Mass, Percent Composition, and the Mole
4.5: Empirical and Molecular Formulas
4.5.1: Practice Problems- Empirical and Molecular Formulas
4: Molecules, Compounds and Quantifying Chemicals is shared under a not declared license and was authored, remixed, and/or curated by
LibreTexts.
1
4.1: Chemical Bonding
Skills to Develop
Define ionic and molecular (covalent) compounds
Predict the type of compound formed from elements based on their location within the periodic table
Determine formulas for simple ionic compounds
Identify polyatomic ions
Atomic Bonding
Chemical bonds form when electrons can be simultaneously close to two or more nuclei, but beyond this, there is no simple, easily
understood theory that would not only explain why atoms bind together to form molecules, but would also predict the three-
dimensional structures of the resulting compounds as well as the energies and other properties of the bonds themselves.
Unfortunately, no one theory exists that accomplishes these goals in a satisfactory way for all of the many categories of compounds
that are known. Moreover, it seems likely that if such a theory does ever come into being, it will be far from simple.
When we are faced with a scientific problem of this complexity, experience has shown that it is often more useful to concentrate
instead on developing models. A scientific model is something like a theory in that it should be able to explain observed
phenomena and to make useful predictions. But whereas a theory can be discredited by a single contradictory case, a model can be
useful even if it does not encompass all instances of the phenomena it attempts to explain.
Given the extraordinary variety of ways in which atoms combine into aggregates, it should come as no surprise that a number of
useful bonding models have been developed. Most of them apply only to certain classes of compounds, or attempt to explain only a
restricted range of phenomena. In this section we will provide brief descriptions of some of the bonding models.
4.1.1 https://chem.libretexts.org/@go/page/217260
ions of opposite charge, and is repelled by ions of like charge; this combination of attractions and repulsions, acting in all
directions, causes the ion to be tightly fixed in its own location in the crystal lattice.
Ionic compounds can often be recognized by their properties. Ionic compounds are solids that typically melt at high temperatures
and boil at even higher temperatures. For example, sodium chloride melts at 801 °C and boils at 1413 °C. (As a comparison, the
molecular compound water melts at 0 °C and boils at 100 °C.) In solid form, an ionic compound is not electrically conductive
because its ions are unable to flow (“electricity” is the flow of charged particles). When molten, however, it can conduct electricity
because its ions are able to move freely through the liquid (Figure 4.1.3; Video 4.1.3).
4.1.2 https://chem.libretexts.org/@go/page/217260
Figure 4.1.2 : Sodium chloride melts at 801 °C and conducts electricity when molten. (credit: modification of work by Mark
Blaser and Matt Evans)
Video 4.1.3 : Watch this video to see a mixture of salts melt and conduct electricity.
In every ionic compound, the total number of positive charges of the cations equals the total number of negative charges of the
anions. Thus, ionic compounds are electrically neutral overall, even though they contain positive and negative ions. We can use this
observation to help us write the formula of an ionic compound. The formula of an ionic compound must have a ratio of ions such
that the numbers of positive and negative charges are equal.
Figure 4.1.4 : Although pure aluminum oxide is colorless, trace amounts of iron and titanium give blue sapphire its
characteristic color. (credit: modification of work by Stanislav Doronenko)
Solution Because the ionic compound must be electrically neutral, it must have the same number of positive and negative
charges. Two aluminum ions, each with a charge of 3+, would give us six positive charges, and three oxide ions, each with a
charge of 2−, would give us six negative charges. The formula would be Al2O3.
4.1.3 https://chem.libretexts.org/@go/page/217260
Exercise 4.1.3
Predict the formula of the ionic compound formed between the sodium cation, Na+, and the sulfide anion, S2−.
Answer
Na2S
Exercise 4.1.2
Using the periodic table, predict whether the following compounds are ionic or covalent:
4.1.4 https://chem.libretexts.org/@go/page/217260
a. SO2
b. CaF2
c. N2H4
d. Al2(SO4)3
Answer a
molecular
Answer b
ionic
Answer c
molecular
Answer d
ionic
Polyatomic Ions
The ions that we have discussed so far are called monatomic ions, that is, they are ions formed from only one atom. Polyatomic
ions are are composed of more than one atom, united chemically. For example, the ammonium ion consists of one nitrogen atoms
and four hydrogen atoms linked together via covalent bonds, and together they comprise a single ion with a 1+ charge and a
formula of NH . The carbonate ion consists of one carbon atom and three oxygen atoms and carries an overall charge of 2−. The
+
formula of the carbonate ion is CO . The atoms of a polyatomic ion are covalently bonded together and so the entire ion behaves
2 −
3
as a single unit.
Some of the more important polyatomic ions are listed in Table . Because you will use them repeatedly, they will soon
4.1.1
become familiar.
Note that there is a system for naming some polyatomic ions: Oxyanions are polyatomic ions that contain one or more oxygen
atoms; -ate and -ite are suffixes designating polyatomic ions containing more or fewer oxygen atoms. Per- (short for “hyper”) and
hypo- (meaning “under”) are prefixes meaning more oxygen atoms than -ate and fewer oxygen atoms than -ite, respectively. For
example, perchlorate is ClO , chlorate is ClO , chlorite is ClO and hypochlorite is ClO−. Unfortunately, the number of oxygen
−
4
−
3
−
2
atoms corresponding to a given suffix or prefix is not consistent; for example, nitrate is NO while sulfate is SO .
−
3
2 −
ammonium NH
+
4
hydronium H O
3
+
oxide O
2 −
peroxide O
2 −
hydroxide OH
−
acetate CH COO
3
−
acetic acid CH COOH
3
cyanide CN
−
hydrocyanic acid HCN
azide N
−
3
hydrazoic acid HN
3
carbonate CO
2 −
3
carbonic acid H CO
2 3
bicarbonate HCO
−
nitrate NO
−
3
nitric acid HNO
3
nitrite NO
−
2
nitrous acid HNO
2
4.1.5 https://chem.libretexts.org/@go/page/217260
Name Formula Related Acid Formula
sulfate SO
2 −
4
sulfuric acid H SO
2 4
sulfite SO
2 −
3
sulfurous acid H SO
2 3
phosphate PO
3 −
4
phosphoric acid H PO
3 4
dihydrogen phosphate H PO
2
−
perchlorate ClO
−
4
perchloric acid HClO
4
chlorate ClO
−
3
chloric acid HClO
3
chlorite ClO
−
2
chlorous acid HClO
2
hypochlorite ClO
−
hypochlorous acid HClO
chromate CrO
2 −
4
chromic acid H CrO
2 4
dichromate Cr O
2
2 −
7
dichromic acid H Cr O
2 2 7
permanganate MnO
−
4
permanganic acid HMnO
4
4
−
3
ammonium carries a 1+ charge and nitrate carries a 1− charge, the two ions will be in a 1:1 ratio and the ionic compound will be
electrically neutral; the formula for ammonium nitrate is NH N O .
4 3
An example of an ionic compound formed between a monoatomic and a polyatomic ion is lithium dichromate: Li C r O . Note 2 2 7
that the lithium cation, Li , carries a 1+ charge and the dichromate anion, Cr O , carries a 2− charge, therefore in order for the
+
2 2
2
7
formula to be neutral, there must be two lithium ions for every one dichromate ion. Use the table of polyatomic ioncs (Table 4.1.1)
when predicting formulas for, or naming ionic compounds that include a polyatomic ion.
designate this by enclosing the formula for the dihydrogen phosphate ion in parentheses and adding a subscript 2. The formula is
Ca(H2PO4)2.
Exercise 4.1.3
Predict the formula of the ionic compound formed between the lithium ion and the peroxide ion, O
2−
2
(Hint: Use the periodic
table to predict the sign and the charge on the lithium ion.)
Answer
Li2O2
4.1.6 https://chem.libretexts.org/@go/page/217260
Summary
The nature of the attractive forces that hold atoms or ions together within a compound is the basis for classifying chemical bonding.
When electrons are transferred and ions form, ionic bonds result. Ionic bonds are electrostatic forces of attraction, that is, the
attractive forces experienced between objects of opposite electrical charge (in this case, cations and anions). When electrons are
“shared” and molecules form, covalent bonds result. Covalent bonds are the attractive forces between the positively charged nuclei
of the bonded atoms and one or more pairs of electrons that are located between the atoms. Compounds are classified as ionic or
molecular (covalent) on the basis of the bonds present in them.
Metals (particularly those in groups 1 and 2) tend to lose the number of electrons that would leave them with the same number of
electrons as in the preceding noble gas in the periodic table. By this means, a positively charged ion is formed. Similarly, nonmetals
(especially those in groups 16 and 17, and, to a lesser extent, those in Group 15) can gain the number of electrons needed to
provide atoms with the same number of electrons as in the next noble gas in the periodic table. Thus, nonmetals tend to form
negative ions. Positively charged ions are called cations, and negatively charged ions are called anions. Ions can be either
monatomic (containing only one atom) or polyatomic (containing more than one atom).
Compounds that contain ions are called ionic compounds. Ionic compounds generally form from metals and nonmetals.
Compounds that do not contain ions, but instead consist of atoms bonded tightly together in molecules (uncharged groups of atoms
that behave as a single unit), are called covalent compounds. Covalent compounds usually form from two or more nonmetals.
Glossary
binary compound
compound containing two different elements.
covalent bond
attractive force between the nuclei of a molecule’s atoms and pairs of electrons between the atoms
covalent compound
(also, molecular compound) composed of molecules formed by atoms of two or more different elements
ionic bond
electrostatic forces of attraction between the oppositely charged ions of an ionic compound
ionic compound
4.1.7 https://chem.libretexts.org/@go/page/217260
compound composed of cations and anions combined in ratios, yielding an electrically neutral substance
molecular compound
(also, covalent compound) composed of molecules formed by atoms of two or more different elements
monatomic ion
ion composed of a single atom
polyatomic ion
ion composed of more than one atom
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
TED-Ed’s commitment to creating lessons worth sharing is an extension of TED’s mission of spreading great ideas. Within
TED-Ed’s growing library of TED-Ed animations, you will find carefully curated educational videos, many of which represent
collaborations between talented educators and animators nominated through the TED-Ed website.
Fuse School, Open Educational Resource free of charge, under a Creative Commons License: Attribution-NonCommercial CC
BY-NC (View License Deed: https://creativecommons.org/licenses/by-nc/4.0/)
Stephen Lower, Professor Emeritus (Simon Fraser U.) Chem1 Virtual Textbook
4.1: Chemical Bonding is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
4.1.8 https://chem.libretexts.org/@go/page/217260
4.2: Representing ionic and molecular compounds and molecules
Skills to Develop
Symbolize the composition of molecules using molecular formulas and empirical formulas
Represent the bonding arrangement of atoms within molecules using structural formulas
Communicating the identity and composition of chemicals is an imperative skill for scientists and there are a variety of methods
that can be used. Each method has a set of rules, and communicates different types of information about the compound or
molecules. This section explains the use of chemical formulas for covalent molecules. The next section will discuss naming, or
nomenclature of molecular and ionic compounds. The use of chemical formulas and nomenclature are used in academic and
industrial settings and can be important for engineers, biologists, geologists, medical professionals and many others.
Two other common representations of molecular (covalent) molecules and compounds are the ball-and-stick model and the space-
filling model. A ball-and-stick model shows the geometric arrangement of the atoms with atomic sizes not to scale, and a space-
filling model shows the relative sizes of the atoms. Figure two shows the molecular and structural formulas for methane and the
ball and stick and space filling models for methane.
4.2.1 https://chem.libretexts.org/@go/page/217261
Figure 4.2.2 : A methane molecule can be represented as (a) a molecular formula, (b) a structural formula, (c) a ball-and-stick
model, and (d) a space-filling model. Carbon and hydrogen atoms are represented by black and white spheres, respectively.
In many cases, the molecular formula of a substance is derived from experimental determination of both its empirical formula and
its molecular mass (the sum of atomic masses for all atoms composing the molecule). For example, it can be determined
experimentally that benzene contains two elements, carbon (C) and hydrogen (H), and that for every carbon atom in benzene, there
is one hydrogen atom. Thus, the empirical formula is CH. An experimental determination of the molecular mass reveals that a
molecule of benzene contains six carbon atoms and six hydrogen atoms, so the molecular formula for benzene is C6H6 (Figure
4.2.3).
Figure 4.2.3 : Benzene, C6H6, is produced during oil refining and has many industrial uses. A benzene molecule can be
represented as (a) a structural formula, (b) a ball-and-stick model, and (c) a space-filling model. (d) Benzene is a clear liquid.
(credit d: modification of work by Sahar Atwa).
If we know a compound’s formula, we can easily determine the empirical formula. (This is somewhat of an academic exercise; the
reverse chronology is generally followed in actual practice.) For example, the molecular formula for acetic acid, the component
that gives vinegar its sharp taste, is C2H4O2. This formula indicates that a molecule of acetic acid (Figure 4.2.4) contains two
carbon atoms, four hydrogen atoms, and two oxygen atoms. The ratio of atoms is 2:4:2. Dividing by the lowest common
denominator (2) gives the simplest, whole-number ratio of atoms, 1:2:1, so the empirical formula is CH2O. Note that a molecular
formula is always a whole-number multiple of an empirical formula.
Figure 4.2.4 : (a) Vinegar contains acetic acid, C2H4O2, which has an empirical formula of CH2O. It can be represented as (b) a
structural formula and (c) as a ball-and-stick model. (credit a: modification of work by “HomeSpot HQ”/Flickr)
4.2.2 https://chem.libretexts.org/@go/page/217261
The molecular formula is C6H12O6 because one molecule actually contains 6 C, 12 H, and 6 O atoms. The simplest whole-
number ratio of C to H to O atoms in glucose is 1:2:1, so the empirical formula is CH2O.
Exercise 4.2.1
A molecule of metaldehyde (a pesticide used for snails and slugs) contains 8 carbon atoms, 16 hydrogen atoms, and 4 oxygen
atoms. What are the molecular and empirical formulas of metaldehyde?
Answer
Molecular formula, C8H16O4; empirical formula, C2H4O
Click to Run
Figure 4.2.5 : Molecules of (a) acetic acid and methyl formate (b) are structural isomers; they have the same formula (C2H4O2)
but different structures (and therefore different chemical properties).
4.2.3 https://chem.libretexts.org/@go/page/217261
Many other types of isomers exist (Figure 4.2.6). Acetic acid and methyl formate are structural isomers, compounds in which the
molecules differ in how the atoms are connected to each other. There are also various types of spatial isomers, in which the relative
orientations of the atoms in space can be different. For example, the compound carvone (found in caraway seeds, spearmint, and
mandarin orange peels) consists of two isomers that are mirror images of each other. S-(+)-carvone smells like caraway, and R-(−)-
carvone smells like spearmint. To learn more about these types of isomers, students can take an organic chemistry course!
Figure 4.2.6 : Molecules of carvone are spatial isomers; they only differ in the relative orientations of the atoms in space. (credit
bottom left: modification of work by “Miansari66”/Wikimedia Commons; credit bottom right: modification of work by Forest &
Kim Starr)
Figure 4.2.3 : A molecule of sulfur is composed of eight sulfur atoms and is therefore written as S8. It can be represented as (a) a
structural formula, (b) a ball-and-stick model, and (c) a space-filling model. Sulfur atoms are represented by yellow spheres.
It is important to note that a subscript following a symbol and a number in front of a symbol do not represent the same thing; for
example, H2 and 2H represent distinctly different species. H2 is a molecular formula; it represents a diatomic molecule of
hydrogen, consisting of two atoms of the element that are chemically bonded together. The expression 2H, on the other hand,
indicates two separate hydrogen atoms that are not combined as a unit. The expression 2H2 represents two molecules of diatomic
hydrogen (Figure 4.2.3).
4.2.4 https://chem.libretexts.org/@go/page/217261
Figure 4.2.3 : The symbols H, 2H, H2, and 2H2 represent very different entities.
Summary
A molecular formula uses chemical symbols and subscripts to indicate the exact numbers of different atoms in a molecule or
compound. An empirical formula gives the simplest, whole-number ratio of atoms in a compound. A structural formula indicates
the bonding arrangement of the atoms in the molecule. Ball-and-stick and space-filling models show the geometric arrangement of
atoms in a molecule. Isomers are compounds with the same molecular formula but different arrangements of atoms.
Going Beyond
What is it that chemists do? According to Lee Cronin, chemists make very complicated molecules by "chopping up" small
molecules and "reverse engineering" them. He wonders if we could "make a really cool universal chemistry set" by what he calls
"app-ing" chemistry. Could we "app" chemistry?
Sorry: we can't play video on this browser. Please make sure it's up to date and that Flash
11.1 or higher is installed. Load this talk on ted.com
Video 4.2.2 : 2012 TED talk about Cronin's idea to "print" medicine.
4.2.5 https://chem.libretexts.org/@go/page/217261
Glossary
empirical formula
formula showing the composition of a compound given as the simplest whole-number ratio of atoms
isomers
compounds with the same chemical formula but different structures
molecular formula
formula indicating the composition of a molecule of a compound and giving the actual number of atoms of each element in a
molecule of the compound.
spatial isomers
compounds in which the relative orientations of the atoms in space differ
structural isomer
one of two substances that have the same molecular formula but different physical and chemical properties because their atoms
are bonded differently
structural formula
shows the atoms in a molecule and how they are connected
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
4.2: Representing ionic and molecular compounds and molecules is shared under a not declared license and was authored, remixed, and/or curated
by LibreTexts.
4.2.6 https://chem.libretexts.org/@go/page/217261
4.2.1: Practice Problems- Writing Chemical Formulas
PROBLEM 4.2.1.1
Explain why the symbol for an atom of the element oxygen and the formula for a molecule of oxygen differ.
Answer
The symbol for the element oxygen, O, represents both the element and one atom of oxygen. A molecule of oxygen, O2,
contains two oxygen atoms; the subscript 2 in the formula must be used to distinguish the diatomic molecule from two single
oxygen atoms.
PROBLEM 4.2.1.2
Write the molecular and empirical formulas of the following compounds:
(a)
(b)
(c)
(d)
Answer a
molecular: CO2
empirical: CO2
Answer b
molecular: C2H2
empirical: CH
Answer c
molecular: C2H4
empirical: CH2
Answer d
molecular: H2SO4
empirical: H2SO4
4.2.1.1 https://chem.libretexts.org/@go/page/217265
PROBLEM 4.2.1.3
Write the molecular and empirical formulas of the following compounds:
(a)
(b)
(c)
(d)
Answer a
molecular: C4H8
empirical: CH2
Answer b
molecular: C4H6
empirical: C2H3
Answer c
molecular: H2Si2Cl4
empirical: HSiCl2
Answer d
molecular: H3PO4
empirical: H3PO4
4.2.1.2 https://chem.libretexts.org/@go/page/217265
Problem 3.3.3
PROBLEM 4.2.1.4
Determine the empirical formulas for the following compounds:
(a) caffeine, C8H10N4O2
(b) fructose, C12H22O11
(c) hydrogen peroxide, H2O2
(d) glucose, C6H12O6
(e) ascorbic acid (vitamin C), C6H8O6
Answer a
C4H5N2O
Answer b
C12H22O11
Answer c
HO
Answer d
CH2O
Answer e
C3H4O3
PROBLEM 4.2.1.5
Determine the empirical formulas for the following compounds:
(a) acetic acid, C2H4O2
(b) citric acid, C6H8O7
(c) hydrazine, N2H4
(d) nicotine, C10H14N2
(e) butane, C4H10
Answer a
CH2O
Answer b
4.2.1.3 https://chem.libretexts.org/@go/page/217265
C6H8O7
Answer c
NH2
Answer d
C5H7N
Answer e
C2H5
Problem 3.3.5
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
4.2.1: Practice Problems- Writing Chemical Formulas is shared under a not declared license and was authored, remixed, and/or curated by
LibreTexts.
4.2.1.4 https://chem.libretexts.org/@go/page/217265
4.3: Chemical Nomenclature
Skills to Develop
Derive names for common types of inorganic compounds using a systematic approach
To name compounds, we need to consider the answers to several questions. First, is the compound ionic or molecular? If the
compound is ionic, does the metal form ions of only one type (fixed charge) or more than one type (variable charge)? Are the ions
monatomic or polyatomic? If the compound is molecular, does it contain hydrogen? If so, does it also contain oxygen? From the
answers we derive, we place the compound in an appropriate category and then name it accordingly. We will begin with the
nomenclature rules for ionic compounds.
4.3.1 https://chem.libretexts.org/@go/page/217262
Compounds Containing Polyatomic Ions
Compounds containing polyatomic ions are named similarly to those containing only monatomic ions, except there is no need to
change to an –ide ending, since the suffix is already present in the name of the anion. Examples are shown in Table 4.3.2.
Table 4.3.2: Names of Some Polyatomic Ionic Compounds
KC2H3O2, potassium acetate (NH4)Cl, ammonium chloride
4.3.2 https://chem.libretexts.org/@go/page/217262
Transition Metal Ionic Compound Name
Out-of-date nomenclature used the suffixes –ic and –ous to designate metals with higher and lower charges, respectively: Iron(III)
chloride, FeCl3, was previously called ferric chloride, and iron(II) chloride, FeCl2, was known as ferrous chloride. Though this
naming convention has been largely abandoned by the scientific community, it remains in use by some segments of industry. For
example, you may see the words stannous fluoride on a tube of toothpaste. This represents the formula SnF2, which is more
properly named tin(II) fluoride. The other fluoride of tin is SnF4, which was previously called stannic fluoride but is now named
tin(IV) fluoride.
neutral. Because the total number of positive charges in each compound must equal the total number of negative charges, the
positive ions must be Fe3+, Cu2+, Ga3+, Cr3+, and Ti3+. These charges are used in the names of the metal ions:
a. iron(III) sulfide
b. copper(II) selenide
c. gallium(III) nitride
d. chromium(III) chloride
e. titanium(III) sulfate
Exercise 4.3.1
Write the formulas of the following ionic compounds:
a. chromium(III) phosphide
b. mercury(II) sulfide
c. manganese(II) phosphate
d. copper(I) oxide
e. chromium(VI) fluoride
Answer a
CrP
Answer b
HgS
Answer c
Mn3(PO4)2
Answer d
Cu2O
Answer e
4.3.3 https://chem.libretexts.org/@go/page/217262
CrF6
Figure 4.3.5 : (a) Erin Brockovich found that Cr(VI), used by PG&E, had contaminated the Hinckley, California, water supply.
(b) The Cr(VI) ion is often present in water as the polyatomic ions chromate, CrO (left), and dichromate, Cr O (right).
2−
4 2
2−
7
Chromium compounds are widely used in industry, such as for chrome plating, in dye-making, as preservatives, and to prevent
corrosion in cooling tower water, as occurred near Hinckley. In the environment, chromium exists primarily in either the Cr(III)
or Cr(VI) forms. Cr(III), an ingredient of many vitamin and nutritional supplements, forms compounds that are not very soluble
in water, and it has low toxicity. Cr(VI), on the other hand, is much more toxic and forms compounds that are reasonably soluble
in water. Exposure to small amounts of Cr(VI) can lead to damage of the respiratory, gastrointestinal, and immune systems, as
well as the kidneys, liver, blood, and skin.
Despite cleanup efforts, Cr(VI) groundwater contamination remains a problem in Hinckley and other locations across the globe.
A 2010 study by the Environmental Working Group found that of 35 US cities tested, 31 had higher levels of Cr(VI) in their tap
water than the public health goal of 0.02 parts per billion set by the California Environmental Protection Agency.
4.3.4 https://chem.libretexts.org/@go/page/217262
Table 4.3.5: Nomenclature Prefixes
N P
u r
m e
Prefix Number
b f
e i
r x
1
(
s
o
m
e
t
i h
m e
e mono- 6 x
s a
o -
m
i
t
t
e
d
)
h
e
p
2 di- 7
t
a
-
o
c
3 tri- 8 t
a
-
n
o
4 tetra- 9 n
a
-
d
e
5 penta- 10 c
a
-
When only one atom of the first element is present, the prefix mono- is usually deleted from that part. Thus, CO is named carbon
monoxide, and CO is called carbon dioxide. When two vowels are adjacent, the a in the Greek prefix is usually dropped. Some
2
4.3.5 https://chem.libretexts.org/@go/page/217262
C
o
m N
p a
Name Compound
o m
u e
n
d
b
o
r
o
n
t
r
S
i
O sulfur dioxide BCl3
c
2
h
l
o
r
i
d
e
s
u
l
f
u
r
h
e
S
x
O sulfur trioxide SF6
a
3
f
l
u
o
r
i
d
e
4.3.6 https://chem.libretexts.org/@go/page/217262
C
o
m N
p a
Name Compound
o m
u e
n
d
p
h
o
s
p
h
o
r
u
s
N p
O nitrogen dioxide PF5 e
2 n
t
a
f
l
u
o
r
i
d
e
t
e
t
r
a
p
h
o
s
p
N h
2 o
dinitrogen tetroxide P4O10
O r
4 u
s
d
e
c
a
o
x
i
d
e
4.3.7 https://chem.libretexts.org/@go/page/217262
C
o
m N
p a
Name Compound
o m
u e
n
d
i
o
d
i
n
e
h
e
N
p
2
dinitrogen pentoxide IF7 t
O
a
5
f
l
u
o
r
i
d
e
There are a few common names that you will encounter as you continue your study of chemistry. For example, although NO is
often called nitric oxide, its proper name is nitrogen monoxide. Similarly, N2O is known as nitrous oxide even though our rules
would specify the name dinitrogen monoxide. (And H2O is usually called water, not dihydrogen monoxide.) You should commit to
memory the common names of compounds as you encounter them.
Exercise 4.3.2
Write the formulas for the following compounds:
a. phosphorus pentachloride
b. dinitrogen monoxide
c. iodine heptafluoride
d. carbon tetrachloride
4.3.8 https://chem.libretexts.org/@go/page/217262
Answer a
PCl5
Answer b
N2O
Answer c
IF7
Answer d
CCl4
Binary Acids
Some compounds containing hydrogen are members of an important class of substances known as acids. The chemistry of these
compounds is explored in more detail in later chapters of this text, but for now, it will suffice to note that many acids release
hydrogen ions, H+, when dissolved in water. To denote this distinct chemical property, a mixture of water with an acid is given a
name derived from the compound’s name. If the compound is a binary acid (comprised of hydrogen and one other nonmetallic
element):
1. The word “hydrogen” is changed to the prefix hydro-
2. The other nonmetallic element name is modified by adding the suffix -ic
3. The word “acid” is added as a second word
For example, when the gas HCl (hydrogen chloride) is dissolved in water, the solution is called hydrochloric acid. Several other
examples of this nomenclature are shown in Table 4.3.7.
Table 4.3.7: Names of Some Simple Acids
Name of Gas Name of Acid
Oxyacids
Many compounds containing three or more elements (such as organic compounds or coordination compounds) are subject to
specialized nomenclature rules that you will learn later. However, we will briefly discuss the important compounds known as
oxyacids, compounds that contain hydrogen, oxygen, and at least one other element, and are bonded in such a way as to impart
acidic properties to the compound (you will learn the details of this in a later chapter). Typical oxyacids consist of hydrogen
combined with a polyatomic, oxygen-containing ion. To name oxyacids:
1. Omit “hydrogen”
2. Start with the root name of the anion
3. Replace –ate with –ic, or –ite with –ous
4. Add “acid”
For example, consider H2CO3 (which you might be tempted to call “hydrogen carbonate”). To name this correctly, “hydrogen” is
omitted; the –ate of carbonate is replace with –ic; and acid is added—so its name is carbonic acid. Other examples are given in
Table 4.3.8. There are some exceptions to the general naming method (e.g., H2SO4 is called sulfuric acid, not sulfic acid, and
H2SO3 is sulfurous, not sulfous, acid).
Table 4.3.8: Names of Common Oxyacids
Formula Anion Name Acid Name
4.3.9 https://chem.libretexts.org/@go/page/217262
Formula Anion Name Acid Name
Summary
Video 4.3.1 : "You know what's an awful lot like unexpectedly waking up in Belguim with a bunch of people speaking several
different languages you don't understand? Chemistry."
Chemists use nomenclature rules to clearly name compounds. Ionic and molecular compounds are named using somewhat-different
methods. Binary ionic compounds typically consist of a metal and a nonmetal. The name of the metal is written first, followed by
the name of the nonmetal with its ending changed to –ide. For example, K2O is called potassium oxide. If the metal can form ions
with different charges, a Roman numeral in parentheses follows the name of the metal to specify its charge. Thus, FeCl2 is iron(II)
chloride and FeCl3 is iron(III) chloride. Some compounds contain polyatomic ions; the names of common polyatomic ions should
be memorized. Molecular compounds can form compounds with different ratios of their elements, so prefixes are used to specify
the numbers of atoms of each element in a molecule of the compound. Examples include SF6, sulfur hexafluoride, and N2O4,
dinitrogen tetroxide. Acids are an important class of compounds containing hydrogen and having special nomenclature rules.
Binary acids are named using the prefix hydro-, changing the –ide suffix to –ic, and adding “acid;” HCl is hydrochloric acid.
Oxyacids are named by changing the ending of the anion (-ate to –ic, and -ite to -ous) and adding “acid;” H2CO3 is carbonic acid.
Glossary
binary acid
compound that contains hydrogen and one other element, bonded in a way that imparts acidic properties to the compound
(ability to release H+ ions when dissolved in water)
binary compound
4.3.10 https://chem.libretexts.org/@go/page/217262
compound containing two different elements.
covalent bond
attractive force between the nuclei of a molecule’s atoms and pairs of electrons between the atoms
covalent compound
(also, molecular compound) composed of molecules formed by atoms of two or more different elements
ionic bond
electrostatic forces of attraction between the oppositely charged ions of an ionic compound
ionic compound
compound composed of cations and anions combined in ratios, yielding an electrically neutral substance
molecular compound
(also, covalent compound) composed of molecules formed by atoms of two or more different elements
monatomic ion
ion composed of a single atom
nomenclature
system of rules for naming objects of interest
oxyacid
compound that contains hydrogen, oxygen, and one other element, bonded in a way that imparts acidic properties to the
compound (ability to release H+ ions when dissolved in water)
oxyanion
polyatomic anion composed of a central atom bonded to oxygen atoms
polyatomic ion
ion composed of more than one atom
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
Stephen Lower, Professor Emeritus (Simon Fraser U.) Chem1 Virtual Textbook
4.3: Chemical Nomenclature is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
4.3.11 https://chem.libretexts.org/@go/page/217262
4.3.1: Practice Problems- Molecular and Ionic Compounds
PROBLEM 4.3.1.1
Using the periodic table, predict whether the following chlorides are ionic or covalent: KCl, NCl3, ICl, MgCl2, PCl5, and CCl4.
Answer
Ionic: KCl, MgCl2; Covalent: NCl3, ICl, PCl5, CCl4
PROBLEM 4.3.1.2
Using the periodic table, predict whether the following chlorides are ionic or covalent: SiCl4, PCl3, CaCl2, CsCl, CuCl2, and
CrCl3.
Answer
Ionic: CaCl2, CuCl2, CsCl, CrCl3; Covalent: SiCl4, PCl3
PROBLEM 4.3.1.3
For each of the following compounds, state whether it is ionic or covalent. If it is ionic, write the symbols for the ions involved:
(a) NF3
(b) BaO
(c) (NH4)2CO3
(d) Sr(H2PO4)2
(e) IBr
(f) Na2O
Answer a
covalent
Answer b
ionic, Ba2+, O2−
Answer c
ionic, NH , CO
+
4
2−
3
Answer d
ionic, Sr2+, H 2
PO
−
Answer e
covalent
Answer f
ionic, Na+, O2−
PROBLEM 4.3.1.4
For each of the following pairs of ions, write the symbol for the formula of the compound they will form:
(a) Ca2+, S2−
(b) NH , SO
+
4
2−
4
(e) Mg2+, PO 3−
4.3.1.1 https://chem.libretexts.org/@go/page/217266
Answer a
CaS
Answer b
(NH4)2CO3
Answer c
AlBr3
Answer d
Na2HPO4
Answer e
Mg3(PO4)2
PROBLEM 4.3.1.5
For each of the following pairs of ions, write the symbol for the formula of the compound they will form:
(a) K+, O2−
(b) NH , PO
+
4
3−
4
(e) Ba2+, PO 3−
Answer a
K2O
Answer b
(NH4)3PO4
Answer c
Al2O3
Answer d
Na2CO3
Answer e
Ba3(PO4)2
4.3.1.2 https://chem.libretexts.org/@go/page/217266
Problem 3.4.5
PROBLEM 4.3.1.6
Name the following compounds:
(a) CsCl
(b) BaO
(c) K2S
(d) BeCl2
(e) HBr
(f) AlF3
Answer a
cesium chloride
Answer b
barium oxide
Answer c
potassium sulfide
Answer d
beryllium chloride
Answer e
hydrogen bromide
Answer f
aluminum fluoride
PROBLEM 4.3.1.7
Name the following compounds:
(a) NaF
(b) Rb2O
(c) BCl3
(d) H2Se
(e) P4O6
(f) ICl3
4.3.1.3 https://chem.libretexts.org/@go/page/217266
Answer a
sodium fluoride
Answer b
Rubidium oxide
Answer c
boron trichloride
Answer d
hydrogen selenide
Answer e
tetraphosphorous hexaoxide
(if you are googling the answers to your homework, google may disagree with you. But we are naming based on the rules we just learned, which
is why you shouldn't trust google for the answers to your homework)
Answer f
Iodine trichloride
PROBLEM 4.3.1.8
Write the formulas of the following compounds:
(a) rubidium bromide
(b) magnesium selenide
(c) sodium oxide
(d) calcium chloride
(e) hydrogen fluoride
(f) gallium phosphide
(g) aluminum bromide
(h) ammonium sulfate
Answer a
RbBr
Answer b
MgSe
Answer c
Na2O
Answer d
CaCl2
Answer e
HF
Answer f
GaP
Answer g
AlBr3
4.3.1.4 https://chem.libretexts.org/@go/page/217266
Answer h
(NH4)2SO4
PROBLEM 4.3.1.9
Write the formulas of the following compounds:
(a) lithium carbonate
(b) sodium perchlorate
(c) barium hydroxide
(d) ammonium carbonate
(e) sulfuric acid
(f) calcium acetate
(g) magnesium phosphate
(h) sodium sulfite
Answer a
Li2CO3
Answer b
NaClO4
Answer c
Ba(OH)2
Answer d
(NH4)2CO3
Answer e
H2SO4
Answer f
Ca(C2H3O2)2
Answer g
Mg3(PO4)2
Answer h
Na2SO3
PROBLEM 4.3.1.10
Write the formulas of the following compounds:
(a) chlorine dioxide
(b) dinitrogen tetraoxide
(c) potassium phosphide
(d) silver(I) sulfide
(e) aluminum nitride
(f) silicon dioxide
Answer a
ClO2
4.3.1.5 https://chem.libretexts.org/@go/page/217266
Answer b
N2O4
Answer c
K3P
Answer d
Ag2S
Answer e
AlN
Answer f
SiO2
PROBLEM 4.3.1.11
Write the formulas of the following compounds:
(a) barium chloride
(b) magnesium nitride
(c) sulfur dioxide
(d) nitrogen trichloride
(e) dinitrogen trioxide
(f) tin(IV) chloride
Answer a
BaCl2
Answer b
Mg3N2
Answer c
SO2
Answer d
NCl3
Answer e
N2O3
Answer f
SnCl4
PROBLEM 4.3.1.12
Each of the following compounds contains a metal that can exhibit more than one ionic charge. Name these compounds:
(a) Cr2O3
(b) FeCl2
(c) CrO3
(d) TiCl4
(e) CoO
(f) MoS2
4.3.1.6 https://chem.libretexts.org/@go/page/217266
Answer a
chromium(III) oxide
Answer b
iron(II) chloride
Answer c
chromium(VI) oxide
Answer d
titanium(IV) chloride
Answer e
cobalt(II) oxide
Answer f
molybdenum(IV) sulfide
PROBLEM 4.3.1.13
Each of the following compounds contains a metal that can exhibit more than one ionic charge. Name these compounds:
(a) NiCO3
(b) MoO3
(c) Co(NO3)2
(d) V2O5
(e) MnO2
(f) Fe2O3
Answer a
nickel (II) carbonate
Answer b
Molybdenum (VI) oxide
Answer c
cobalt (II) nitrate
Answer d
vanadium (V) oxide
Answer e
manganese (IV) oxide
Answer f
iron (III) oxide
PROBLEM 4.3.1.14
The following ionic compounds are found in common household products. Write the formulas for each compound:
(a) potassium phosphate
(b) copper(II) sulfate
(c) calcium chloride
4.3.1.7 https://chem.libretexts.org/@go/page/217266
(d) titanium dioxide
(e) ammonium nitrate
(f) sodium bisulfate (the common name for sodium hydrogen sulfate)
Answer a
K3PO4
Answer b
CuSO4
Answer c
CaCl2
Answer d
TiO2
Answer e
NH4NO3
Answer f
NaHSO4
PROBLEM 4.3.1.15
The following ionic compounds are found in common household products. Name each of the compounds:
(a) Ca(H2PO4)2
(b) FeSO4
(c) CaCO3
(d) MgO
(e) NaNO2
(f) KI
Answer a
calcium dihydrogen phosphate
Answer b
iron (II) sulfate
Answer c
calcium carbonate
Answer d
magnesium oxide
Answer e
sodium nitrite
Answer f
potassium iodide
4.3.1.8 https://chem.libretexts.org/@go/page/217266
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
4.3.1: Practice Problems- Molecular and Ionic Compounds is shared under a not declared license and was authored, remixed, and/or curated by
LibreTexts.
4.3.1.9 https://chem.libretexts.org/@go/page/217266
4.4: Formula Mass, Percent Composition, and the Mole
Skills to Develop
Calculate formula masses for covalent and ionic compounds
Define the amount unit mole and the related quantity Avogadro’s number
Explain the relation between mass, moles, and numbers of atoms or molecules, and perform calculations deriving these
quantities from one another
Compute the percent composition of a compound
We can argue that modern chemical science began when scientists started exploring the quantitative as well as the qualitative
aspects of chemistry. For example, Dalton’s atomic theory was an attempt to explain the results of measurements that allowed him
to calculate the relative masses of elements combined in various compounds. Understanding the relationship between the masses of
atoms and the chemical formulas of compounds allows us to quantitatively describe the composition of substances.
Formula Mass
In an earlier chapter, we described the development of the atomic mass unit, the concept of average atomic masses, and the use of
chemical formulas to represent the elemental makeup of substances. These ideas can be extended to calculate the formula mass of a
substance by summing the average atomic masses of all the atoms represented in the substance’s formula.
Figure 4.4.1 : The average mass of a chloroform molecule, CHCl3, is 119.37 amu, which is the sum of the average atomic masses
of each of its constituent atoms. The model shows the molecular structure of chloroform.
Likewise, the molecular mass of an aspirin molecule, C9H8O4, is the sum of the atomic masses of nine carbon atoms, eight
hydrogen atoms, and four oxygen atoms, which amounts to 180.15 amu (Figure 4.4.2).
Figure 4.4.2 : The average mass of an aspirin molecule is 180.15 amu. The model shows the molecular structure of aspirin,
C9H8O4.
4.4.1 https://chem.libretexts.org/@go/page/217263
Ibuprofen, C13H18O2, is a covalent compound and the active ingredient in several popular nonprescription pain medications,
such as Advil and Motrin. What is the molecular mass (amu) for this compound?
Solution
Molecules of this compound are comprised of 13 carbon atoms, 18 hydrogen atoms, and 2 oxygen atoms. Following the
approach described above, the average molecular mass for this compound is therefore:
Exercise 4.4.1
Acetaminophen, C8H9NO2, is a covalent compound and the active ingredient in several popular nonprescription pain
medications, such as Tylenol. What is the molecular mass (amu) for this compound?
Answer
151.16 amu
Figure 4.4.3 : Table salt, NaCl, contains an array of sodium and chloride ions combined in a 1:1 ratio. Its formula mass is 58.44
amu.
Note that the average masses of neutral sodium and chlorine atoms were used in this computation, rather than the masses for
sodium cations and chlorine anions. This approach is perfectly acceptable when computing the formula mass of an ionic
compound. Even though a sodium cation has a slightly smaller mass than a sodium atom (since it is missing an electron), this
difference will be offset by the fact that a chloride anion is slightly more massive than a chloride atom (due to the extra electron).
Moreover, the mass of an electron is negligibly small with respect to the mass of a typical atom. Even when calculating the mass of
an isolated ion, the missing or additional electrons can generally be ignored, since their contribution to the overall mass is
negligible, reflected only in the nonsignificant digits that will be lost when the computed mass is properly rounded. The few
exceptions to this guideline are very light ions derived from elements with precisely known atomic masses.
4.4.2 https://chem.libretexts.org/@go/page/217263
Aluminum sulfate, Al2(SO4)3, is an ionic compound that is used in the manufacture of paper and in various water purification
processes. What is the formula mass (amu) of this compound?
Solution
The formula for this compound indicates it contains Al3+ and SO42− ions combined in a 2:3 ratio. For purposes of computing a
formula mass, it is helpful to rewrite the formula in the simpler format, Al2S3O12. Following the approach outlined above, the
formula mass for this compound is calculated as follows:
Exercise 4.4.2
Calcium phosphate, Ca3(PO4)2, is an ionic compound and a common anti-caking agent added to food products. What is the
formula mass (amu) of calcium phosphate?
Answer
310.18 amu
The Mole
In units 1 and 2, we learned about measuring quanity of atoms using the mole. As a reminder, the mole is an amount unit similar to
familiar units like pair, dozen, gross, etc. The number of entities composing a mole has been experimentally determined to be
6.02214179 × 10 , a fundamental constant named Avogadro’s number (NA) or the Avogadro constant in honor of Italian scientist
23
Amedeo Avogadro. This constant is properly reported with an explicit unit of “per mole,” a conveniently rounded version being
/mol.
23
6.022 × 10
The relationships between formula mass, the mole, and Avogadro’s number can be applied to compute various quantities that
describe the composition of substances and compounds. In unit 2, we used Avogadro’s number to convert between mass of an
element, moles of an element, and number of atoms in the sample. Now, we can use Avogadro’s number and formula mass (also
called molecular weight) to convent between mass of a sample, moles, and number of molecules. For example, if we know the
mass and chemical composition of a substance, we can determine the number of moles and calculate number of atoms or molecules
in the sample. Likewise, if we know the number of moles of a substance, we can derive the number of atoms or molecules and
calculate the substance’s mass.s
4.4.3 https://chem.libretexts.org/@go/page/217263
The molar mass of glycine is required for this calculation, and it is computed in the same fashion as its molecular mass. One
mole of glycine, C2H5O2N, contains 2 moles of carbon, 5 moles of hydrogen, 2 moles of oxygen, and 1 mole of nitrogen:
The provided mass of glycine (~28 g) is a bit more than one-third the molar mass (~75 g/mol), so we would expect the
computed result to be a bit greater than one-third of a mole (~0.33 mol). Dividing the compound’s mass by its molar mass
yields:
mol glycine
28.35 g glycine ( ) = 0.378 mol glycine
75.07 g
Exercise 4.4.7
How many moles of sucrose, C 12 H22 O11 , are in a 25-g sample of sucrose?
Answer
0.073 mol
The molar mass for this compound is computed to be 176.124 g/mol. The given number of moles is a very small fraction of a
mole (~10−4 or one-ten thousandth); therefore, we would expect the corresponding mass to be about one-ten thousandth of the
molar mass (~0.02 g). Performing the calculation, we get:
176.124 g
−4
1.42 × 10 mol vitamin C ( ) = 0.0250 g vitamin C
mol vitamin C
Exercise 4.4.8
What is the mass of 0.443 mol of hydrazine, N 2 H4 ?
Answer
14.2 g
Example 4.4.9 : Deriving the Number of Molecules from the Compound Mass
4.4.4 https://chem.libretexts.org/@go/page/217263
A packet of an artificial sweetener contains 40.0 mg of saccharin (C7H5NO3S), which has the structural formula:
Given that saccharin has a molar mass of 183.18 g/mol, how many saccharin molecules are in a 40.0-mg (0.0400-g) sample of
saccharin? How many carbon atoms are in the same sample?
Solution
The number of molecules in a given mass of compound is computed by first deriving the number of moles, as demonstrated in
Example 4.4.8, and then multiplying by Avogadro’s number:
Using the provided mass and molar mass for saccharin yields:
23
mol C H NO S 6.022 × 10 C H NO S molecules
7 5 3 7 5 3
0.0400 g C H NO S ( )( ) (4.4.1)
7 5 3
183.18 g C H NO S
7 5 3 1 mol C H NO S
7 5 3
20
= 1.31 × 10 C H NO S molecules
7 5 3
The compound’s formula shows that each molecule contains seven carbon atoms, and so the number of C atoms in the provided
sample is:
20
7 C atoms 21
1.31 × 10 C H NO S molecules ( ) = 9.20 × 10 C atoms
7 5 3
1 C H NO S molecule
7 5 3
Exercise 4.4.9
How many C 4 H10 molecules are contained in 9.213 g of this compound? How many hydrogen atoms?
Answer
22
9.545 × 10 molecules C4 H10
23
9.545 × 10 atoms H
4.4.5 https://chem.libretexts.org/@go/page/217263
Video 4.4.3 : A preview of some of the uses we will have for moles in upcoming units
Percent Composition
In the previous section, we discussed the relationship between the bulk mass of a substance and the number of atoms or molecules
it contains (moles). Given the chemical formula of the substance, we were able to determine the amount of the substance (moles)
from its mass, and vice versa. But what if the chemical formula of a substance is unknown? In this section, we will explore how to
apply these very same principles in order to derive the chemical formulas of unknown substances from experimental mass
measurements.
The elemental makeup of a compound defines its chemical identity, and chemical formulas are the most succinct way of
representing this elemental makeup. When a compound’s formula is unknown, measuring the mass of each of its constituent
elements is often the first step in the process of determining the formula experimentally. The results of these measurements permit
the calculation of the compound’s percent composition, defined as the percentage by mass of each element in the compound. For
example, consider a gaseous compound composed solely of carbon and hydrogen. The percent composition of this compound could
be represented as follows:
mass H
%H = × 100% (4.4.2)
mass compound
mass C
%C = × 100% (4.4.3)
mass compound
If analysis of a 10.0-g sample of this gas showed it to contain 2.5 g H and 7.5 g C, the percent composition would be calculated to
be 25% H and 75% C:
2.5 g H
%H = × 100% = 25% (4.4.4)
10.0 g compound
7.5 g C
%C = × 100% = 75% (4.4.5)
10.0 g compound
1.85 g H
%H = × 100% = 15.4%
12.04 g compound
2.85 g N
%N = × 100% = 23.7%
12.04 g compound
The analysis results indicate that the compound is 61.0% C, 15.4% H, and 23.7% N by mass.
Exercise 4.4.10
A 24.81-g sample of a gaseous compound containing only carbon, oxygen, and chlorine is determined to contain 3.01 g C, 4.00
g O, and 17.81 g Cl. What is this compound’s percent composition?
Answer
12.1% C, 16.1% O, 71.8% Cl
4.4.6 https://chem.libretexts.org/@go/page/217263
Determining Percent Composition from Formula Mass
Video 4.4.4 : A video overview of how to calculate percent composition of a compound based on its chemical formula.
Percent composition is also useful for evaluating the relative abundance of a given element in different compounds of known
formulas. As one example, consider the common nitrogen-containing fertilizers ammonia (NH3), ammonium nitrate (NH4NO3),
and urea (CH4N2O). The element nitrogen is the active ingredient for agricultural purposes, so the mass percentage of nitrogen in
the compound is a practical and economic concern for consumers choosing among these fertilizers. For these sorts of applications,
the percent composition of a compound is easily derived from its formula mass and the atomic masses of its constituent elements.
A molecule of NH3 contains one N atom weighing 14.01 amu and three H atoms weighing a total of (3 × 1.008 amu) = 3.024 amu.
The formula mass of ammonia is therefore (14.01 amu + 3.024 amu) = 17.03 amu, and its percent composition is:
14.01 amu N
%N = × 100% = 82.27% (4.4.6)
17.03 amu NH3
3.024 amu N
%H = × 100% = 17.76% (4.4.7)
17.03 amu NH3
This same approach may be taken considering a pair of molecules, a dozen molecules, or a mole of molecules, etc. The latter
amount is most convenient and would simply involve the use of molar masses instead of atomic and formula masses, as
demonstrated Example 4.4.2. As long as we know the chemical formula of the substance in question, we can easily derive percent
composition from the formula mass or molar mass.
%C = 60.00 % C
%H = 4.476 % H
%O = 35.52%
4.4.7 https://chem.libretexts.org/@go/page/217263
Note that these percentages sum to equal 100.00% when appropriately rounded.
Exercise 4.4.11
To three significant digits, what is the mass percentage of iron in the compound F e 2 O3 ?
Answer
69.9% Fe
Summary
The formula mass of a substance is the sum of the average atomic masses of each atom represented in the chemical formula and is
expressed in atomic mass units. The formula mass of a covalent compound is also called the molecular mass. A convenient amount
unit for expressing very large numbers of atoms or molecules is the mole. Experimental measurements have determined the number
of entities composing 1 mole of substance to be 6.022 × 1023, a quantity called Avogadro’s number. The mass in grams of 1 mole of
substance is its molar mass. Due to the use of the same reference substance in defining the atomic mass unit and the mole, the
formula mass (amu) and molar mass (g/mol) for any substance are numerically equivalent (for example, one H2O molecule weighs
approximately 18 amu and 1 mole of H2O molecules weighs approximately 18 g).
Footnotes
1. 1 Omiatek, Donna M., Amanda J. Bressler, Ann-Sofie Cans, Anne M. Andrews, Michael L. Heien, and Andrew G. Ewing. “The
Real Catecholamine Content of Secretory Vesicles in the CNS Revealed by Electrochemical Cytometry.” Scientific Report 3
(2013): 1447, accessed January 14, 2015, doi:10.1038/srep01447.
2. Read more about the redefinition of SI units including the kilogram here (Laura Howe, CE&N, Nov. 16, 2018).
Glossary
Avogadro’s number (NA)
experimentally determined value of the number of entities comprising 1 mole of substance, equal to 6.022 × 1023 mol−1
formula mass
sum of the average masses for all atoms represented in a chemical formula; for covalent compounds, this is also the molecular
mass
mole
amount of substance containing the same number of atoms, molecules, ions, or other entities as the number of atoms in exactly
12 grams of 12C
molar mass
mass in grams of 1 mole of a substance
percent composition
percentage by mass of the various elements in a compound
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
TED-Ed’s commitment to creating lessons worth sharing is an extension of TED’s mission of spreading great ideas. Within
TED-Ed’s growing library of TED-Ed animations, you will find carefully curated educational videos, many of which represent
collaborations between talented educators and animators nominated through the TED-Ed website.
4.4.8 https://chem.libretexts.org/@go/page/217263
Fuse School, Open Educational Resource free of charge, under a Creative Commons License: Attribution-NonCommercial CC
BY-NC (View License Deed: https://creativecommons.org/licenses/by-nc/4.0/)
4.4: Formula Mass, Percent Composition, and the Mole is shared under a not declared license and was authored, remixed, and/or curated by
LibreTexts.
4.4.9 https://chem.libretexts.org/@go/page/217263
4.4.1: Practice Problems- Formula Mass, Percent Composition, and the Mole
PROBLEM 4.4.1.1
What is the total mass (amu) of carbon in each of the following molecules?
(a) CH4
(b) CHCl3
(c) C12H10O6
(d) CH3CH2CH2CH2CH3
Answer a
12.01 amu
Answer b
12.01 amu
Answer c
144.12 amu
Answer d
60.05 amu
Problem 4.2.1
PROBLEM 4.4.1.2
What is the total mass of hydrogen in each of the molecules?
(a) CH4
(b) CHCl3
(c) C12H10O6
(d) CH3CH2CH2CH2CH3
Answer a
4.4.1.1 https://chem.libretexts.org/@go/page/217267
4.032 amu
Answer b
1.008 amu
Answer c
10.08 amu
Answer d
12.096 amu
PROBLEM 4.4.1.3
Calculate the molecular or formula mass of each of the following:
(a) P4
(b) H2O
(c) Ca(NO3)2
(d) CH3CO2H (acetic acid)
(e) C12H22O11 (sucrose, cane sugar).
Answer a
123.896 amu
Answer b
18.015 amu
Answer c
164.086 amu
Answer d
60.052 amu
Answer e
342.297 amu
4.4.1.2 https://chem.libretexts.org/@go/page/217267
Problem 4.2.3
PROBLEM 4.4.1.4
Determine the molecular mass of the following compounds:
(a)
(b)
(c)
(d)
Answer a
98.906 amu
Answer b
26.018 amu
4.4.1.3 https://chem.libretexts.org/@go/page/217267
Answer c
185.818 amu
Answer d
98.072 amu
PROBLEM 4.4.1.5
Determine the molecular mass of the following compounds:
(a)
(b)
(c)
(d)
Answer a
56.107 amu
Answer b
54.091 amu
Answer c
199.9976 amu
Answer d
97.9950 amu
4.4.1.4 https://chem.libretexts.org/@go/page/217267
Problem 4.2.5
PROBLEM 4.4.1.6
Compare 1 mole of H2, 1 mole of O2, and 1 mole of F2.
(a) Which has the largest number of molecules? Explain why.
(b) Which has the greatest mass? Explain why.
Answer a
1 mole is always 6.022 x 1023 molecules. They have the same number of molecules.
Answer b
F2; it has the highest molar mass
PROBLEM 4.4.1.7
Which contains the greatest mass of oxygen: 0.75 mol of ethanol (C2H5OH), 0.60 mol of formic acid (HCO2H), or 1.0 mol of
water (H2O)? Explain why.
Answer
Formic acid. Its formula has twice as many oxygen atoms as the other two compounds (one each). Therefore, 0.60 mol of
formic acid would be equivalent to 1.20 mol of a compound containing a single oxygen atom.
PROBLEM 4.4.1.8
Calculate the molar mass of each of the following:
(a) S8
(b) C5H12
(c) Sc2(SO4)3
(d) CH3COCH3 (acetone)
(e) C6H12O6 (glucose)
Answer a
4.4.1.5 https://chem.libretexts.org/@go/page/217267
256.528 g/mol
Answer b
72.150 g/mol
Answer c
378.103 g/mol
Answer d
58.080 g/mol
Answer e
180.158 g/mol
PROBLEM 4.4.1.9
Calculate the molar mass of each of the following:
(a) the anesthetic halothane, C2HBrClF3
(b) the herbicide paraquat, C12H14N2Cl2
(c) caffeine, C8H10N4O2
(d) urea, CO(NH2)2
(e) a typical soap, C17H35CO2Na
Answer a
197.382 g/mol
Answer b
257.163 g/mol
Answer c
194.193 g/mol
Answer d
60.056 g/mol
Answer e
306.464 g/mol
4.4.1.6 https://chem.libretexts.org/@go/page/217267
Problem 4.2.9
PROBLEM 4.4.1.10
Determine the number of moles of compound and the number of moles of each type of atom in each of the following:
(a) 25.0 g of propylene, C3H6
(b) 3.06 × 10−3 g of the amino acid glycine, C2H5NO2
(c) 25 lb of the herbicide Treflan, C13H16N2O4F (1 lb = 454 g)
(d) 0.125 kg of the insecticide Paris Green, Cu4(AsO3)2(CH3CO2)2
(e) 325 mg of aspirin, C6H4(CO2H)(CO2CH3)
Answer a
0.595 mol C3H6
1.78 mol C
3.57 mol H
Answer b
4.08 × 10-5 mol C2H5NO2
8.16 × 10-5 mol C
2.04 × 10-4 mol H
4.08 × 10-5 mol N
8.16 × 10-5 mol O
Answer c
40.106 mol C13H16N2O4F
521.37 mol C
641.70 mol H
80.212 mol N
160.42 mol O
4.4.1.7 https://chem.libretexts.org/@go/page/217267
40.106 mol F
Answer d
0.202 mol Cu4(AsO3)2(CH3CO2)2
0.808 mol Cu
0.404 mol As
2.02 mol O
0.808 mol C
1.212 mol H
Answer e
0.00181 mol C6H4(CO2H)(CO2CH3)
0.01625 mol C
0.0144 mol H
0.0072 mol O
PROBLEM 4.4.1.11
Determine the mass of each of the following:
(a) 0.0146 mol KOH
(b) 10.2 mol ethane, C2H6
(c) 1.6 × 10−3 mol Na2 SO4
(d) 6.854 × 103 mol glucose, C6 H12 O6
(e) 2.86 mol Co(NH3)6Cl3
Answer a
0.819 g
Answer b
307 g
Answer c
0.23 g
Answer d
1.235 × 106 g (1235 kg)
Answer e
765 g
4.4.1.8 https://chem.libretexts.org/@go/page/217267
Problem 4.2.11
PROBLEM 4.4.1.12
Determine the number of moles of the compound and determine the number of moles of each type of atom in each of the
following:
(a) 2.12 g of potassium bromide, KBr
(b) 0.1488 g of phosphoric acid, H3PO4
(c) 23 kg of calcium carbonate, CaCO3
(d) 78.452 g of aluminum sulfate, Al2(SO4)3
(e) 0.1250 mg of caffeine, C8H10N4O2
Answer a
0.0178 mol KBr
0.0178 mol K
0.0178 mol Br
Answer b
0.00152 mol H3PO4
0.00455 mol H
0.00152 mol P
0.00607 mol O
Answer c
230 mol CaCO3
230 mol Ca
230 mol C
690 mol O
Answer d
0.229 mol Al2(SO4)3
4.4.1.9 https://chem.libretexts.org/@go/page/217267
0.459 mol Al
0.688 mol S
2.75 mol O
Answer e
6.44 × 10-7 mol C8H10N4O2
5.15 × 10-6 mol C
6.44 × 10-6 mol H
2.58 × 10-6 mol N
1.29 × 10-6 mol O
PROBLEM 4.4.1.13
The approximate minimum daily dietary requirement of the amino acid leucine, C6H13NO2, is 1.1 g. What is this requirement in
moles?
Answer
0.0084 mol C6H13NO2
Problem 4.2.13
PROBLEM 4.4.1.14
Determine the mass in grams of each of the following:
(a) 0.600 mol of oxygen atoms
(b) 0.600 mol of oxygen molecules, O2
(c) 0.600 mol of ozone molecules, O3
Answer a
9.60 g
Answer b
19.2 g
Answer c
4.4.1.10 https://chem.libretexts.org/@go/page/217267
28.2 g
PROBLEM 4.4.1.15
A 55-kg woman has 7.5 × 10−3 mol of hemoglobin (molar mass = 64,456 g/mol) in her blood. How many hemoglobin
molecules is this? What is this quantity in grams?
Answer
4.52 × 1021 molecules of hemoglobin
483.42 g of hemoglobin
Problem 4.2.15
PROBLEM 4.4.1.16
One 55-gram serving of a particular cereal supplies 270 mg of sodium, 11% of the recommended daily allowance. How many
moles and atoms of sodium are in the recommended daily allowance?
Answer
0.107 mol Na; 6.45 × 1022 atoms Na
PROBLEM 4.4.1.17
A tube of toothpaste contains 0.76 g of sodium monofluorophosphate (Na2PO3F) in 100 mL.
a. What mass of fluorine atoms in mg was present?
b. How many fluorine atoms were present?
Answer a
100.27 mg F
Answer b
3.18 × 1021 atoms F
4.4.1.11 https://chem.libretexts.org/@go/page/217267
Problem 4.2.17
PROBLEM 4.4.1.18
Which of the following represents the least number of molecules?
a. 20.0 g of H2O (18.02 g/mol)
b. 77.0 g of CH4 (16.06 g/mol)
c. 68.0 g of CaH2 (42.09 g/mol)
d. 100.0 g of N2O (44.02 g/mol)
e. 84.0 g of HF (20.01 g/mol)
Answer
20.0 g of H2O represents the smallest number of moles, meaning the least number of molecules present. Since 1 mole =
6.022 × 1023 molecules (or atoms) regardless of identity, the least number of moles will equal the least number of molecules
PROBLEM 4.4.1.19
Calculate the following:
(a) the percent composition of ammonia, NH3
(b) the percent composition of photographic “hypo,” Na2S2O3
(c) the percent of calcium ion in Ca3(PO4)2
Answer a
% N = 82.24%
% H = 17.76%
Answer b
% Na = 29.08%
% S = 40.56%
% O = 30.36%
Answer c
% Ca2+ = 38.76%
4.4.1.12 https://chem.libretexts.org/@go/page/217267
Problem 4.2.19
PROBLEM 4.4.1.20
A compound of carbon and hydrogen contains 92.3% C and has a molar mass of 78.1 g/mol. What is its molecular formula?
Answer
C6H6
PROBLEM 4.4.1.21
Dichloroethane, a compound that is often used for dry cleaning, contains carbon, hydrogen, and chlorine. It has a molar mass of
99 g/mol. Analysis of a sample shows that it contains 24.3% carbon and 4.1% hydrogen. What is its molecular formula?
Answer
C2H4Cl2
Problem 4.3.4
In the second edition, this problem was moved. The number in the title of the video may be incorrect, but the solution is
correct.
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
4.4.1.13 https://chem.libretexts.org/@go/page/217267
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
4.4.1: Practice Problems- Formula Mass, Percent Composition, and the Mole is shared under a not declared license and was authored, remixed,
and/or curated by LibreTexts.
4.4.1.14 https://chem.libretexts.org/@go/page/217267
4.5: Empirical and Molecular Formulas
Skills to Develop
Determine the empirical formula of a compound
Determine the molecular formula of a compound
1 mol H
0.287 g H × = 0.284 mol H (4.5.2)
1.008 g H
Thus, we can accurately represent this compound with the formula C0.142H0.284. Of course, per accepted convention, formulas
contain whole-number subscripts, which can be achieved by dividing each subscript by the smaller subscript:
C 0.142 H 0.284 or CH (4.5.3)
2
0.142 0.142
(Recall that subscripts of “1” are not written, but rather assumed if no other number is present.)
The empirical formula for this compound is thus CH2. This may or not be the compound’s molecular formula as well; however, we
would need additional information to make that determination (as discussed later in this section).
Consider as another example a sample of compound determined to contain 5.31 g Cl and 8.40 g O. Following the same approach
yields a tentative empirical formula of:
C l0.150O0.525 = C l 0.150 O 0.525 = Cl O3.5 (4.5.4)
0.150 0.150
In this case, dividing by the smallest subscript still leaves us with a decimal subscript in the empirical formula. To convert this into
a whole number, we must multiply each of the subscripts by two, retaining the same atom ratio and yielding Cl2O7 as the final
empirical formula.
Procedure
In summary, empirical formulas are derived from experimentally measured element masses by:
1. Deriving the number of moles of each element from its mass
2. Dividing each element’s molar amount by the smallest molar amount to yield subscripts for a tentative empirical formula
3. Multiplying all coefficients by an integer, if necessary, to ensure that the smallest whole-number ratio of subscripts is
obtained
Figure 4.5.1 outlines this procedure in flow chart fashion for a substance containing elements A and X.
4.5.1 https://chem.libretexts.org/@go/page/217264
Figure 4.5.1 : The empirical formula of a compound can be derived from the masses of all elements in the sample.
Figure 4.5.2 : Hematite is an iron oxide that is used in jewelry. (credit: Mauro Cateb)
Solution
For this problem, we are given the mass in grams of each element. Begin by finding the moles of each:
mol Fe
34.97 g Fe ( ) = 0.6261 mol Fe
55.85 g
mol O
15.03 g O ( ) = 0.9394 mol O
16.00 g
Next, derive the iron-to-oxygen molar ratio by dividing by the lesser number of moles:
0.6261
= 1.000 mol Fe
0.6261
0.9394
= 1.500 mol O
0.6261
The ratio is 1.000 mol of iron to 1.500 mol of oxygen (Fe1O1.5). Finally, multiply the ratio by two to get the smallest possible
whole number subscripts while still maintaining the correct iron-to-oxygen ratio:
2(F e1 O1.5 ) = F e2 O3
Exercise 4.5.3
What is the empirical formula of a compound if a sample contains 0.130 g of nitrogen and 0.370 g of oxygen?
Answer
N2 O5
4.5.2 https://chem.libretexts.org/@go/page/217264
Calculating Percent Composition and E…
E…
Video 4.5.2 : Additional worked examples illustrating the derivation of empirical formulas are presented in the brief video clip.
Figure 4.5.3 : An oxide of carbon is removed from these fermentation tanks through the large copper pipes at the top. (credit:
“Dual Freq”/Wikimedia Commons)
Solution
Since the scale for percentages is 100, it is most convenient to calculate the mass of elements present in a sample weighing 100
g. The calculation is “most convenient” because, per the definition for percent composition, the mass of a given element in
grams is numerically equivalent to the element’s mass percentage. This numerical equivalence results from the definition of the
“percentage” unit, whose name is derived from the Latin phrase per centum meaning “by the hundred.” Considering this
definition, the mass percentages provided may be more conveniently expressed as fractions:
27.29 g C
27.29 % C =
100 g compound
72.71 g O
72.71 % O =
100 g compound
The molar amounts of carbon and hydrogen in a 100-g sample are calculated by dividing each element’s mass by its molar mass:
4.5.3 https://chem.libretexts.org/@go/page/217264
mol C
27.29 g C ( ) = 2.272 mol C
12.01 g
mol O
72.71 g O ( ) = 4.544 mol O
16.00 g
Coefficients for the tentative empirical formula are derived by dividing each molar amount by the lesser of the two:
2.272 mol C
=1
2.272
4.544 mol O
=2
2.272
Since the resulting ratio is one carbon to two oxygen atoms, the empirical formula is CO2.
Exercise 4.5.4
What is the empirical formula of a compound containing 40.0% C, 6.71% H, and 53.28% O?
Answer
C H2 O
Video 4.5.3 : A review of calculating empirical formula from percent composition and an explanation of deriving molecular
formula.
Recall that empirical formulas are symbols representing the relative numbers of a compound’s elements. Determining the absolute
numbers of atoms that compose a single molecule of a covalent compound requires knowledge of both its empirical formula and its
molecular mass or molar mass. These quantities may be determined experimentally by various measurement techniques. Molecular
mass, for example, is often derived from the mass spectrum of the compound (see discussion of this technique in the previous
chapter on atoms and molecules). Molar mass can be measured by a number of experimental methods, many of which will be
introduced in later chapters of this text.
Molecular formulas are derived by comparing the compound’s molecular or molar mass to its empirical formula mass. As the name
suggests, an empirical formula mass is the sum of the average atomic masses of all the atoms represented in an empirical formula.
If we know the molecular (or molar) mass of the substance, we can divide this by the empirical formula mass in order to identify
the number of empirical formula units per molecule, which we designate as n:
g
molecular or molar mass (amu or )
mol
= n f ormula units/molecule (4.5.5)
g
empirical f ormula mass (amu or )
mol
4.5.4 https://chem.libretexts.org/@go/page/217264
The molecular formula is then obtained by multiplying each subscript in the empirical formula by n, as shown by the generic
empirical formula AxBy:
(Ax By )n = Anx Bnx (4.5.6)
For example, consider a covalent compound whose empirical formula is determined to be CH2O. The empirical formula mass for
this compound is approximately 30 amu (the sum of 12 amu for one C atom, 2 amu for two H atoms, and 16 amu for one O atom).
If the compound’s molecular mass is determined to be 180 amu, this indicates that molecules of this compound contain six times
the number of atoms represented in the empirical formula:
180 amu/molecule
= 6 f ormula units/molecule (4.5.7)
amu
30
f ormula unit
Molecules of this compound are then represented by molecular formulas whose subscripts are six times greater than those in the
empirical formula:
(CH O) =C H O (4.5.8)
2 6 6 12 6
Note that this same approach may be used when the molar mass (g/mol) instead of the molecular mass (amu) is used. In this case,
we are merely considering one mole of empirical formula units and molecules, as opposed to single units and molecules.
1 mol H
(8.710 g H) ( ) = 8.624 mol H (4.5.10)
1.01 g H
1 mol N
(17.27 g N) ( ) = 1.233 mol N (4.5.11)
14.01 g N
Next, we calculate the molar ratios of these elements relative to the least abundant element, N .
6.163 mol C / 1.233 mol N = 5
8.264 mol H / 1.233 mol N = 7
1.233 mol N / 1.233 mol N = 1
1.233/1.233 = 1.000 mol N
6.163/1.233 = 4.998 mol C
8.624/1.233 = 6.994 mol H
The C-to-N and H-to-N molar ratios are adequately close to whole numbers, and so the empirical formula is C5H7N. The
empirical formula mass for this compound is therefore 81.13 amu/formula unit, or 81.13 g/mol formula unit.
We calculate the molar mass for nicotine from the given mass and molar amount of compound:
40.57 g nicotine 162.3 g
=
0.2500 mol nicotine mol
Comparing the molar mass and empirical formula mass indicates that each nicotine molecule contains two formula units:
4.5.5 https://chem.libretexts.org/@go/page/217264
162.3 g/mol
= 2 f ormula units/molecule
g
81.13
f ormula unit
Thus, we can derive the molecular formula for nicotine from the empirical formula by multiplying each subscript by two:
(C H N) =C H N
5 7 2 10 14 2
Exercise 4.5.5
What is the molecular formula of a compound with a percent composition of 49.47% C, 5.201% H, 28.84% N, and 16.48% O,
and a molecular mass of 194.2 amu?
Answer
C8H10N4O2
Summary
The chemical identity of a substance is defined by the types and relative numbers of atoms composing its fundamental entities
(molecules in the case of covalent compounds, ions in the case of ionic compounds). A compound’s percent composition provides
the mass percentage of each element in the compound, and it is often experimentally determined and used to derive the compound’s
empirical formula. The empirical formula mass of a covalent compound may be compared to the compound’s molecular or molar
mass to derive a molecular formula.
Key Equations
mass X
%X = × 100%
mass compound
g
molecular or molar mass (amu or )
mol
= n f ormula units/molecule
g
empirical f ormula mass (amu or )
mol
(AxBy)n = AnxBny
Glossary
percent composition
percentage by mass of the various elements in a compound
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Fuse School, Open Educational Resource free of charge, under a Creative Commons License: Attribution-NonCommercial CC
BY-NC (View License Deed: https://creativecommons.org/licenses/by-nc/4.0/)
4.5: Empirical and Molecular Formulas is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
4.5.6 https://chem.libretexts.org/@go/page/217264
4.5.1: Practice Problems- Empirical and Molecular Formulas
PROBLEM 4.5.1.1
Determine the empirical formulas for compounds with the following percent compositions:
(a) 15.8% carbon and 84.2% sulfur
(b) 40.0% carbon, 6.7% hydrogen, and 53.3% oxygen
Answer a
CS2
Answer b
CH2O
Problem 4.3.1
PROBLEM 4.5.1.2
Determine the empirical and molecular formula for chrysotile asbestos. Chrysotile has the following percent composition:
28.03% Mg, 21.60% Si, 1.16% H, and 49.21% O. The molar mass for chrysotile is 520.8 g/mol.
Answer
Mg3Si2H3O8 (empirical formula), Mg6Si4H6O16 (molecular formula)
PROBLEM 4.5.1.3
Polymers are large molecules composed of simple units repeated many times. Thus, they often have relatively simple empirical
formulas. Calculate the empirical formulas of the following polymers:
(a) Lucite (Plexiglas); 59.9% C, 8.06% H, 32.0% O
(b) Saran; 24.8% C, 2.0% H, 73.1% Cl
(c) polyethylene; 86% C, 14% H
(d) polystyrene; 92.3% C, 7.7% H
(e) Orlon; 67.9% C, 5.70% H, 26.4% N
4.5.1.1 https://chem.libretexts.org/@go/page/217268
Answer a
C5H8O2
Answer b
CHCl
Answer c
CH2
Answer d
CH
Answer e
C3H3N
Problem 4.3.3
PROBLEM 4.5.1.4
A major textile dye manufacturer developed a new yellow dye. The dye has a percent composition of 75.95% C, 17.72% N, and
6.33% H by mass with a molar mass of about 240 g/mol. Determine the molecular formula of the dye.
Answer
C15H15N3
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
4.5.1: Practice Problems- Empirical and Molecular Formulas is shared under a not declared license and was authored, remixed, and/or curated by
LibreTexts.
4.5.1.2 https://chem.libretexts.org/@go/page/217268
CHAPTER OVERVIEW
5: Transformations of Matter
Unit 5 Objectives
By the end of this unit, you will be able to:
Derive chemical equations from narrative descriptions of chemical reactions.
Write and balance chemical equations
Explain the concept of stoichiometry as it pertains to chemical reactions
Use balanced chemical equations to derive stoichiometric factors relating amounts of reactants and products
Perform stoichiometric calculations involving mass and moles
Explain the concepts of theoretical yield and limiting reactants/reagents.
Derive the theoretical yield for a reaction under specified conditions.
Calculate the percent yield for a reaction.
5: Transformations of Matter is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
1
5.1: Writing and Balancing Chemical Equations
Skills to Develop
Derive chemical equations from narrative descriptions of chemical reactions.
Write and balance chemical equations
Figure 5.1.1 : The reaction between methane and oxygen to yield carbon dioxide and water (shown at bottom) may be represented
by a chemical equation using formulas (top).
This example illustrates the fundamental aspects of any chemical equation:
1. The substances undergoing reaction are called reactants, and their formulas are placed on the left side of the equation.
2. The substances generated by the reaction are called products, and their formulas are placed on the right sight of the equation.
3. Plus signs (+) separate individual reactant and product formulas, and an arrow (⟶) separates the reactant and product (left and
right) sides of the equation.
4. The relative numbers of reactant and product species are represented by coefficients (numbers placed immediately to the left of
each formula). A coefficient of 1 is typically omitted.
It is common practice to use the smallest possible whole-number coefficients in a chemical equation, as is done in this example.
Realize, however, that these coefficients represent the relative numbers of reactants and products, and, therefore, they may be
correctly interpreted as ratios. Methane and oxygen react to yield carbon dioxide and water in a 1:2:1:2 ratio. This ratio is satisfied
5.1.1 https://chem.libretexts.org/@go/page/217270
if the numbers of these molecules are, respectively, 1-2-1-2, or 2-4-2-4, or 3-6-3-6, and so on (Figure 5.1.2). Likewise, these
coefficients may be interpreted with regard to any amount (number) unit, and so this equation may be correctly read in many ways,
including:
One methane molecule and two oxygen molecules react to yield one carbon dioxide molecule and two water molecules.
One dozen methane molecules and two dozen oxygen molecules react to yield one dozen carbon dioxide molecules and two
dozen water molecules.
One mole of methane molecules and 2 moles of oxygen molecules react to yield 1 mole of carbon dioxide molecules and 2
moles of water molecules.
Figure 5.1.2 : Regardless of the absolute number of molecules involved, the ratios between numbers of molecules of each species
that react (the reactants) and molecules of each species that form (the products) are the same and are given in the chemical
equation.
Balancing Equations
⎛ 2 O atoms ⎞ ⎛ 1 O atom ⎞
1 CO molecule × + 2 H O molecule × = 4 O atoms (5.1.1)
2 2
⎝ CO molecule ⎠ ⎝ H O molecule ⎠
2 2
5.1.2 https://chem.libretexts.org/@go/page/217270
The equation for the reaction between methane and oxygen to yield carbon dioxide and water is confirmed to be balanced per this
approach, as shown here:
CH +2 O → CO +2 H O (5.1.2)
4 2 2 2
O 2×2=4 (1 × 2) + (2 × 1) = 4 4 = 4, yes
A balanced chemical equation often may be derived from a qualitative description of some chemical reaction by a fairly simple
approach known as balancing by inspection. Consider as an example the decomposition of water to yield molecular hydrogen and
oxygen. This process is represented qualitatively by an unbalanced chemical equation:
H O→ H +O (unbalanced)
2 2 2
Comparing the number of H and O atoms on either side of this equation confirms its imbalance:
O 1×1=1 1×2=2 1 ≠ 2, no
The numbers of H atoms on the reactant and product sides of the equation are equal, but the numbers of O atoms are not. To
achieve balance, the coefficients of the equation may be changed as needed. Keep in mind, of course, that the formula subscripts
define, in part, the identity of the substance, and so these cannot be changed without altering the qualitative meaning of the
equation. For example, changing the reactant formula from H2O to H2O2 would yield balance in the number of atoms, but doing so
also changes the reactant’s identity (it’s now hydrogen peroxide and not water). The O atom balance may be achieved by changing
the coefficient for H2O to 2.
2H O→ H +O (unbalanced)
2 2 2
H 2×2=4 1×2=2 4 ≠ 2, no
The H atom balance was upset by this change, but it is easily reestablished by changing the coefficient for the H2 product to 2.
2H O→ 2H +O (balanced)
2 2 2
These coefficients yield equal numbers of both H and O atoms on the reactant and product sides, and the balanced equation is,
therefore:
2H O→ 2H +O (5.1.3)
2 2 2
5.1.3 https://chem.libretexts.org/@go/page/217270
How To Balance Equations - Part 2 | Che…
Che…
Video 5.1.3 : A review of balancing chemical equations, with extra practice problems.
Next, count the number of each type of atom present in the unbalanced equation.
O 1×2=2 1×5=5 2 ≠ 5, no
Though nitrogen is balanced, changes in coefficients are needed to balance the number of oxygen atoms. To balance the number
of oxygen atoms, a reasonable first attempt would be to change the coefficients for the O2 and N2O5 to integers that will yield
10 O atoms (the least common multiple for the O atom subscripts in these two formulas).
N +5 O → 2N O (unbalanced)
2 2 2 5
N 1×2=2 2×2=4 2 ≠ 4, no
O 5 × 2 = 10 2 × 5 = 10 10 = 10, yes
The N atom balance has been upset by this change; it is restored by changing the coefficient for the reactant N2 to 2.
2N +5 O → 2N O (5.1.4)
2 2 2 5
O 5 × 2 = 10 2 × 5 = 10 10 = 10, yes
The numbers of N and O atoms on either side of the equation are now equal, and so the equation is balanced.
Exercise 5.1.1
Write a balanced equation for the decomposition of ammonium nitrate to form molecular nitrogen, molecular oxygen, and water.
(Hint: Balance oxygen last, since it is present in more than one molecule on the right side of the equation.)
5.1.4 https://chem.libretexts.org/@go/page/217270
Answer
2 NH NO → 2N +O +4 H O
4 3 2 2 2
It is sometimes convenient to use fractions instead of integers as intermediate coefficients in the process of balancing a chemical
equation. When balance is achieved, all the equation’s coefficients may then be multiplied by a whole number to convert the
fractional coefficients to integers without upsetting the atom balance. For example, consider the reaction of ethane (C2H6) with
oxygen to yield H2O and CO2, represented by the unbalanced equation:
C H +O → H O + CO (unbalanced)
2 6 2 2 2
Following the usual inspection approach, one might first balance C and H atoms by changing the coefficients for the two product
species, as shown:
C H +O → 3 H O + 2 CO (unbalanced)
2 6 2 2 2
This results in seven O atoms on the product side of the equation, an odd number—no integer coefficient can be used with the O2
reactant to yield an odd number, so a fractional coefficient, , is used instead to yield a provisional balanced equation:
7
7
C H + O → 3 H O + 2 CO
2 6 2 2 2 2
A conventional balanced equation with integer-only coefficients is derived by multiplying each coefficient by 2:
2C H +7 O → 6 H O + 4 CO (5.1.5)
2 6 2 2 2
Finally with regard to balanced equations, recall that convention dictates use of the smallest whole-number coefficients. Although
the equation for the reaction between molecular nitrogen and molecular hydrogen to produce ammonia is, indeed, balanced,
3N +9 H → 6 NH (5.1.6)
2 2 3
the coefficients are not the smallest possible integers representing the relative numbers of reactant and product molecules. Dividing
each coefficient by the greatest common factor, 3, gives the preferred equation:
N +3 H → 2 NH (5.1.7)
2 2 3
Phet Simulation
Use this interactive tutorial for additional practice balancing equations.
5.1.5 https://chem.libretexts.org/@go/page/217270
Balancing Chemical Eq
X+Y XY
Game
Introduction
Video 5.1.4 : How to use state symbol to give more information in your chemical equations.
5.1.6 https://chem.libretexts.org/@go/page/217270
The physical states of reactants and products in chemical equations very often are indicated with a parenthetical abbreviation
following the formulas. Common abbreviations include s for solids, l for liquids, g for gases, and aq for substances dissolved in
water (aqueous solutions, as introduced in the preceding chapter). These notations are illustrated in the example equation here:
2 Na(s) + 2 H O(l) → 2 NaOH(aq) + H (g) (5.1.8)
2 2
This equation represents the reaction that takes place when sodium metal is placed in water. The solid sodium reacts with liquid
water to produce molecular hydrogen gas and the ionic compound sodium hydroxide (a solid in pure form, but readily dissolved in
water).
Special conditions necessary for a reaction are sometimes designated by writing a word or symbol above or below the equation’s
arrow. For example, a reaction carried out by heating may be indicated by the uppercase Greek letter delta (Δ) over the arrow.
Δ
CaCO (s) −
−→ CaO(s) + CO (g) (5.1.9)
3 2
Other examples of these special conditions will be encountered in more depth in later units.
Glossary
balanced equation
chemical equation with equal numbers of atoms for each element in the reactant and product
chemical equation
symbolic representation of a chemical reaction
coefficient
number placed in front of symbols or formulas in a chemical equation to indicate their relative amount
molecular equation
chemical equation in which all reactants and products are represented as neutral substances
product
substance formed by a chemical or physical change; shown on the right side of the arrow in a chemical equation
5.1.7 https://chem.libretexts.org/@go/page/217270
reactant
substance undergoing a chemical or physical change; shown on the left side of the arrow in a chemical equation
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Instititute of Technology
Fuse School, Open Educational Resource free of charge, under a Creative Commons License: Attribution-NonCommercial CC
BY-NC (View License Deed: https://creativecommons.org/licenses/by-nc/4.0/)
Crash Course Chemistry, References and image licenses for this episode can be found in the Google document here:
http://dft.ba/-51j; Crash Course is a division of Complexly and videos are free to stream for educational purposes.
5.1: Writing and Balancing Chemical Equations is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
5.1.8 https://chem.libretexts.org/@go/page/217270
5.1.1: Practice Problems- Writing and Balancing Chemical Equations
PROBLEM 5.1.1.1
Balance the following equations:
a. PCl (s) + H O(l) → POCl (l) + HCl(aq)
5 2 3
Answer a
PCl (s) + H O(l) → POCl (l) + 2 HCl(aq)
5 2 3
Answer b
3 Cu(s) + 8 HNO (aq) → 3 Cu(NO ) (aq) + 4 H O(l) + 2 NO(g)
3 3 2 2
Answer c
H (g) + I (s) → 2 HI(s)
2 2
Answer d
4 Fe(s) + 3 O (g) → 2 Fe O (s)
2 2 3
Answer e
2 Na(s) + 2 H O(l) → 2 NaOH(aq) + H (g)
2 2
Answer f
(NH ) Cr O (s) → Cr O (s) + N (g) + 4 H O(g)
4 2 52 7 2 3 2 2
Answer g
P (s) + 6 Cl (g) → 4 PCl (l)
4 2 3
Answer h
PtCl (s) → Pt(s) + 2 Cl (g)
4 2
Problem 5.1.1
5.1.1.1 https://chem.libretexts.org/@go/page/217273
PROBLEM 5.1.1.2
Balance the following equations:
a. Ag(s) + H S(g) + O (g) → Ag S(s) + H O(l)
2 2 2 2
Answer a
4 Ag(s) + 2 H S(g) + O (g) → 2 Ag S(s) + 2 H O(l)
2 2 2 2
Answer b
P (s) + 5 O (g) → P O (s)
4 2 4 10
Answer c
2 Pb(s) + 2 H O(l) + O (g) → 2 Pb(OH) (s)
2 2 2
Answer d
3 Fe(s) + 4 H O(l) → Fe O (s) + 4 H (g)
2 3 4 2
Answer e
Sc O (s) + 3 SO (l) → Sc (SO ) (s)
2 3 3 2 4 3
Answer f
Ca (PO ) (aq) + 4 H PO (aq) → 3 Ca (H PO ) (aq)
3 4 2 3 4 2 4 2
Answer g
2 Al(s) + 3 H SO (aq) → Al (SO ) (s) + 3 H (g)
2 4 2 4 3 2
Answer h
TiCl (s) + 2 H O(g) → TiO (s) + 4 HCl(g)
4 2 2
PROBLEM 5.1.1.3
Write a balanced molecular equation describing each of the following chemical reactions.
a. Solid calcium carbonate is heated and decomposes to solid calcium oxide and carbon dioxide gas.
b. Gaseous butane, C4H10, reacts with diatomic oxygen gas to yield gaseous carbon dioxide and water vapor.
c. Aqueous solutions of magnesium chloride and sodium hydroxide react to produce solid magnesium hydroxide and aqueous
sodium chloride.
d. Water vapor reacts with sodium metal to produce solid sodium hydroxide and hydrogen gas.
Answer a
CaCO (s) → CaO(s) + CO (g)
3 2
Answer b
5.1.1.2 https://chem.libretexts.org/@go/page/217273
2C H (g) + 13 O (g) → 8 CO (g) + 10 H O(g)
4 10 2 2 2
Answer c
MgCl (aq) + 2 NaOH(aq) → Mg (OH) (s) + 2 NaCl(aq)
2 2
Answer d
2 H O(g) + 2 Na(s) → 2 NaOH(s) + H (g)
2 2
Problem 5.1.3
PROBLEM 5.1.1.4
Write a balanced equation describing each of the following chemical reactions.
a. Solid potassium chlorate, KClO3, decomposes to form solid potassium chloride and diatomic oxygen gas.
b. Solid aluminum metal reacts with solid diatomic iodine to form solid Al2I6.
c. When solid sodium chloride is added to aqueous sulfuric acid, hydrogen chloride gas and aqueous sodium sulfate are
produced.
d. Aqueous solutions of phosphoric acid and potassium hydroxide react to produce aqueous potassium dihydrogen phosphate
and liquid water.
Answer a
2 KClO (s) → 2 KCl(s) + 3 O (g)
3 2
Answer b
2 Al(s) + 3 I (s) → Al I (s)
2 2 6
Answer c
2 NaCl(s) + H SO (aq) → 2 HCl(g) + Na SO (aq)
2 4 2 4
Answer d
H PO (aq) + KOH(aq) → KH PO (aq) + H O(l)
3 4 2 4 2
PROBLEM 5.1.1.5
5.1.1.3 https://chem.libretexts.org/@go/page/217273
Colorful fireworks often involve the decomposition of barium nitrate and potassium chlorate and the reaction of the metals
magnesium, aluminum, and iron with oxygen.
a. Write the formulas of barium nitrate and potassium chlorate.
b. The decomposition of solid potassium chlorate leads to the formation of solid potassium chloride and diatomic oxygen gas.
Write an equation for the reaction.
c. The decomposition of solid barium nitrate leads to the formation of solid barium oxide, diatomic nitrogen gas, and diatomic
oxygen gas. Write an equation for the reaction.
d. Write separate equations for the reactions of the solid metals magnesium, aluminum, and iron with diatomic oxygen gas to
yield the corresponding metal oxides. (Assume the iron oxide contains Fe3+ ions.)
Answer a
Ba(NO3)2, KClO3
Answer b
2 KClO (s) → 2 KCl(s) + 3 O (g)
3 2
Answer c
2 Ba (NO ) (s) → 2 BaO(s) + 2 N (g) + 5 O (g)
3 2 2 2
Answer d
2 Mg(s) + O (g) → 2 MgO(s)
2
; 4 Al(s) + 3 O 2
(g) → 2 Al O (g)
2 3
; 4 Fe(s) + 3 O
2
(g) → 2 Fe O (s)
2 3
Problem 5.1.5
PROBLEM 5.1.1.6
Aqueous hydrogen fluoride (hydrofluoric acid) is used to etch glass and to analyze minerals for their silicon content. Hydrogen
fluoride will also react with sand (silicon dioxide).
a. Write an equation for the reaction of solid silicon dioxide with hydrofluoric acid to yield gaseous silicon tetrafluoride and
liquid water.
b. The mineral fluorite (calcium fluoride) occurs extensively in Illinois. Solid calcium fluoride can also be prepared by the
reaction of aqueous solutions of calcium chloride and sodium fluoride, yielding aqueous sodium chloride as the other
product. Write the equation for this reaction.
Answer a
4 HF(aq) + SiO (s) → SiF (g) + 2 H O(l)
2 4 2
5.1.1.4 https://chem.libretexts.org/@go/page/217273
Answer b
CaCl (aq) + 2 NaF(aq) → 2 NaCl(aq) + CaF (s)
2 2
PROBLEM 5.1.1.7
A novel process for obtaining magnesium from sea water involves several reactions. Write a balanced chemical equation for
each step of the process.
a. The first step is the decomposition of solid calcium carbonate from seashells to form solid calcium oxide and gaseous carbon
dioxide.
b. The second step is the formation of solid calcium hydroxide as the only product from the reaction of the solid calcium oxide
with liquid water.
c. Solid calcium hydroxide is then added to the seawater, reacting with dissolved magnesium chloride to yield solid magnesium
hydroxide and aqueous calcium chloride.
d. The solid magnesium hydroxide is added to a hydrochloric acid solution, producing dissolved magnesium chloride and
liquid water.
e. Finally, the magnesium chloride is melted and electrolyzed to yield liquid magnesium metal and diatomic chlorine gas.
Answer a
CaCO (s) → CaO(s) + CO (g)
3 2
Answer b
CaO(s) + H O(l) → Ca (OH) (s)
2 2
Answer c
Ca (OH) (s) + MgCl (aq) → Mg (OH) (s) + CaCl (aq)
2 2 2 2
Answer d
Mg (OH) (s) + 2 HCl(aq) → MgCl (aq) + 2 H O(l)
2 2 2
Answer e
MgCl (s) → Mg(s) + Cl (g)
2 2
Problem 5.1.7
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
5.1.1.5 https://chem.libretexts.org/@go/page/217273
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
5.1.1: Practice Problems- Writing and Balancing Chemical Equations is shared under a not declared license and was authored, remixed, and/or
curated by LibreTexts.
5.1.1.6 https://chem.libretexts.org/@go/page/217273
5.2: Reaction Stoichiometry
Skills to Develop
Explain the concept of stoichiometry as it pertains to chemical reactions
Use balanced chemical equations to derive stoichiometric factors relating amounts of reactants and products
Perform stoichiometric calculations involving mass and moles
3
1 cup mix + cup milk + 1 egg → 8 pancakes (5.2.1)
4
If two dozen pancakes are needed for a big family breakfast, the ingredient amounts must be increased proportionally according to
the amounts given in the recipe. For example, the number of eggs required to make 24 pancakes is
1 egg
24 pancakes × = 3 eggs (5.2.2)
8 pancakes
Balanced chemical equations are used in much the same fashion to determine the amount of one reactant required to react with a
given amount of another reactant, or to yield a given amount of product, and so forth. The coefficients in the balanced equation are
used to derive stoichiometric factors that permit computation of the desired quantity. To illustrate this idea, consider the production
of ammonia by reaction of hydrogen and nitrogen:
N (g) + 3 H (g) → 2 NH (g) (5.2.3)
2 2 3
This equation shows that ammonia molecules are produced from hydrogen molecules in a 2:3 ratio, and stoichiometric factors may
be derived using any amount (number) unit:
2 NH molecules 2 doz NH molecules 2 mol NH molecules
3 3 3
or or (5.2.4)
3H molecules 3 doz H molecules 3 mol H molecules
2 2 2
5.2.1 https://chem.libretexts.org/@go/page/217271
These stoichiometric factors can be used to compute the number of ammonia molecules produced from a given number of
hydrogen molecules, or the number of hydrogen molecules required to produce a given number of ammonia molecules. Similar
factors may be derived for any pair of substances in any chemical equation.
Figure 5.2.1 : Aluminum and iodine react to produce aluminum iodide. The heat of the reaction vaporizes some of the solid
iodine as a purple vapor. (credit: modification of work by Mark Ott)
Solution
3 mol I
Referring to the balanced chemical equation, the stoichiometric factor relating the two substances of interest is 2
. The
2 mol Al
molar amount of iodine is derived by multiplying the provided molar amount of aluminum by this factor:
3 mol I2
mol I2 = 0.429 mol Al ×
2 mol Al
= 0.644 mol I2
Exercise 5.2.1
How many moles of Ca(OH)2 are required to react with 1.36 mol of H3PO4 to produce Ca3(PO4)2 according to the equation
3 Ca (OH) + 2 H PO → Ca (PO ) + 6 H O ?
2 3 4 3 4 2 2
Answer
2.04 mol
Solution
The approach here is the same as for Example 5.2.1, though the absolute number of molecules is requested, not the number of
moles of molecules. This will simply require use of the moles-to-numbers conversion factor, Avogadro’s number.
The balanced equation shows that carbon dioxide is produced from propane in a 3:1 ratio:
3 mol CO
2
(5.2.7)
1 mol C H
3 8
Using this stoichiometric factor, the provided molar amount of propane, and Avogadro’s number,
5.2.2 https://chem.libretexts.org/@go/page/217271
3 mol CO2 23
6.022 × 10 C O2 molecules 24
0.75 mol C3 H8 × × = 1.4 × 10 C O2 molecules (5.2.8)
1 mol C3 H8 mol CO2
Exercise 5.2.1
How many NH3 molecules are produced by the reaction of 4.0 mol of Ca(OH)2 according to the following equation:
Answer
4.8 × 1024 NH3 molecules
These examples illustrate the ease with which the amounts of substances involved in a chemical reaction of known stoichiometry
may be related. Directly measuring numbers of atoms and molecules is, however, not an easy task, and the practical application of
stoichiometry requires that we use the more readily measured property of mass.
Solution
The approach used previously in Examples 5.2.1 and 5.2.2 is likewise used here; that is, we must derive an appropriate
stoichiometric factor from the balanced chemical equation and use it to relate the amounts of the two substances of interest. In
this case, however, masses (not molar amounts) are provided and requested, so additional steps of the sort learned in the
previous chapter are required. The calculations required are outlined in this flowchart:
Exercise 5.2.3
What mass of gallium oxide, Ga2O3, can be prepared from 29.0 g of gallium metal? The equation for the reaction is
4 Ga + 3 O → 2 Ga O .
2 2 3
Answer
39.0 g
5.2.3 https://chem.libretexts.org/@go/page/217271
Example 5.2.4 : Relating Masses of Reactants
What mass of oxygen gas, O2, from the air is consumed in the combustion of 702 g of octane, C8H18, one of the principal
components of gasoline?
2C H + 25 O → 16 CO + 18 H O
8 18 2 2 2
Solution
The approach required here is the same as for the Example 5.2.3 , differing only in that the provided and requested masses are
both for reactant species.
1 mol C H 25 mol O
8 18 2 32.00 g O
2 3
702 g C H × × × = 2.46 × 10 g O
8 18 2
114.23 g C H 2 mol C H mol O
8 18 8 18 2
Exercise 5.2.4
What mass of CO is required to react with 25.13 g of Fe2O3 according to the equation Fe 2
O
3
+ 3 CO → 2 Fe + 3 CO
2
?
Answer
13.22 g
These examples illustrate just a few instances of reaction stoichiometry calculations. Numerous variations on the beginning and
ending computational steps are possible depending upon what particular quantities are provided and sought (volumes, solution
concentrations, and so forth). Regardless of the details, all these calculations share a common essential component: the use of
stoichiometric factors derived from balanced chemical equations. Figure 5.2.2 provides a general outline of the various
computational steps associated with many reaction stoichiometry calculations.
5.2.4 https://chem.libretexts.org/@go/page/217271
Figure 5.2.2 : The flowchart depicts the various computational steps involved in most reaction stoichiometry calculations. We will
discuss Molarity in Unit 6.
Airbags
Airbags (Figure 5.2.3) are a safety feature provided in most automobiles since the 1990s. The effective operation of an airbag
requires that it be rapidly inflated with an appropriate amount (volume) of gas when the vehicle is involved in a collision. This
requirement is satisfied in many automotive airbag systems through use of explosive chemical reactions, one common choice
being the decomposition of sodium azide, NaN3. When sensors in the vehicle detect a collision, an electrical current is passed
through a carefully measured amount of NaN3 to initiate its decomposition:
This reaction is very rapid, generating gaseous nitrogen that can deploy and fully inflate a typical airbag in a fraction of a second
(~0.03–0.1 s). Among many engineering considerations, the amount of sodium azide used must be appropriate for generating
enough nitrogen gas to fully inflate the air bag and ensure its proper function. For example, a small mass (~100 g) of NaN3 will
generate approximately 50 L of N2.
5.2.5 https://chem.libretexts.org/@go/page/217271
Figure 5.2.3 : Airbags deploy upon impact to minimize serious injuries to passengers. (credit: Jon Seidman)
Summary
Video 5.2.2 : An overview of the process covered in this section with a preview of the next section.
A balanced chemical equation may be used to describe a reaction’s stoichiometry (the relationships between amounts of reactants
and products). Coefficients from the equation are used to derive stoichiometric factors that subsequently may be used for
computations relating reactant and product masses, molar amounts, and other quantitative properties.
Glossary
stoichiometric factor
ratio of coefficients in a balanced chemical equation, used in computations relating amounts of reactants and products
stoichiometry
relationships between the amounts of reactants and products of a chemical reaction
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Fuse School, Open Educational Resource free of charge, under a Creative Commons License: Attribution-NonCommercial CC
BY-NC (View License Deed: https://creativecommons.org/licenses/by-nc/4.0/)
5.2: Reaction Stoichiometry is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
5.2.6 https://chem.libretexts.org/@go/page/217271
5.2.1: Practice Problems- Reaction Stoichiometry
PROBLEM 5.2.1.1
Write the balanced equation and determine the information requested. Don't worry about state symbols in these reactions.
a. The number of moles and the mass (in grams) of chlorine, Cl2, required to react with 10.0 g of sodium metal, Na, to produce
sodium chloride, NaCl.
b. The number of moles and the mass (in milligrams) of diatomic oxygen formed by the decomposition of 1.252 g of
mercury(II) oxide.
c. The number of moles and the mass (in g) of sodium nitrate, NaNO3, required to decompose and produce 128 g of diatomic
oxygen, where NaNO2 is the other product.
d. The number of moles and the mass (in kg) of carbon dioxide formed by the combustion of 20.0 kg of carbon in an excess of
diatomic oxygen.
e. The number of moles and the mass (in kg) of copper(II) carbonate needed to decompose in order to produce 1.500 kg of
copper(II) oxide, where CO2 is the other product.
f. The number of moles and mass (in grams) of C2H4 required to react with water to produce 9.55 g C2H6O.
Answer a
2 Na + Cl → 2 NaCl
2
Answer b
2 HgO → 2 Hg + O
2
0.00289 mol O2
92 mg O2
Answer c
2 NaNO → 2 NaNO +O
3 3 2
8 mol NaNO3
680 g NaNO3
Answer d
C +O → CO
2 2
Answer e
CuCO → CuO + CO
3 2
Answer f
C H +H O → C H O
2 4 2 2 6
5.2.1.1 https://chem.libretexts.org/@go/page/217275
*Apologies for the brief phone ringing*
Problem 5.2.1
PROBLEM 5.2.1.2
I2 is produced by the reaction of 0.4235 mol of CuCl2 according to the following equation:
2 CuCl
2
+ 4 KI → 2 CuI + 4 KCl + I
2
.
a. How many molecules of I2 are produced?
b. What mass of I2 is produced?
Answer a
1.28 × 1023 molecules I2
Answer b
53.8 g I2
PROBLEM 5.2.1.3
Silver is often extracted from ores as K[Ag(CN)2] and then recovered by the reaction
2 K[Ag (CN) ](aq) + Zn(s) → 2 Ag(s) + Zn(CN) (aq) + 2 KCN(aq)
2 2
a. How many molecules of Zn(CN)2 are produced by the reaction of 35.27 g of K[Ag(CN)2]?
b. What mass of Zn(CN)2 is produced?
Answer a
5.337 × 1022 molecules
Answer b
10.41 g Zn(CN)2
5.2.1.2 https://chem.libretexts.org/@go/page/217275
Problem 5.2.3
PROBLEM 5.2.1.4
What mass of silver oxide, Ag2O, is required to produce 25.0 g of silver sulfadiazine, AgC10H9N4SO2, from the reaction of
silver oxide and sulfadiazine?
2C H N SO + Ag O → 2 AgC H N SO +H O
10 10 4 2 2 10 9 4 2 2
Answer
8.12 g Ag2O
PROBLEM 5.2.1.5
Carborundum is silicon carbide, SiC, a very hard material used as an abrasive on sandpaper and in other applications. It is
prepared by the reaction of pure sand, SiO2, with carbon at high temperature. Carbon monoxide, CO, is the other product of this
reaction. Write the balanced equation for the reaction, and calculate how much SiO2 is required to produce 3.00 kg of SiC.
Answer
SiO + 3 C → SiC + 2 CO
2
4.50 kg SiO2
Problem 5.2.5
5.2.1.3 https://chem.libretexts.org/@go/page/217275
PROBLEM 5.2.1.6
Automotive air bags inflate when a sample of sodium azide, NaN3, is very rapidly decomposed.
2 NaN (s) → 2 Na(s) + 3 N (g)
3 2
What mass of sodium azide is required to produce 2.6 ft3 (73.6 L) of nitrogen gas with a density of 1.25 g/L?
Answer
142g NaN3
PROBLEM 5.2.1.7
Urea, CO(NH2)2, is manufactured on a large scale for use in producing urea-formaldehyde plastics and as a fertilizer. What is
the maximum mass of urea that can be manufactured from the CO2 produced by combustion of 1.00×103 kg of carbon followed
by the reaction?
CO (g) + 2 NH (g) → CO(NH ) (s) + H O(l) (5.2.1.1)
2 3 2 2 2
Answer
5.00 kg Urea
Problem 5.2.7
PROBLEM 5.2.1.8
In an accident, a solution containing 2.5 kg of nitric acid was spilled. Two kilograms of Na2CO3 was quickly spread on the area
and CO2 was released by the reaction. Was sufficient Na2CO3 used to neutralize all of the acid? (in this reaction, water and
sodium nitrate are the other two products)
Answer
No, you will need 2.1 kg of sodium carbonate to neutralize 2.5 kg of nitric acid.
PROBLEM 5.2.1.9
A compact car gets 37.5 miles per gallon on the highway. If gasoline contains 84.2% carbon by mass and has a density of
0.8205 g/mL, determine the mass of carbon dioxide produced during a 500-mile trip (3.785 liters per gallon).
Answer
5.2.1.4 https://chem.libretexts.org/@go/page/217275
1.28 × 105 g CO2
Click here to see a video of the solution
Problem 5.2.9
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
5.2.1: Practice Problems- Reaction Stoichiometry is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
5.2.1.5 https://chem.libretexts.org/@go/page/217275
5.3: Calculating Reaction Yields
Skills to Develop
Explain the concepts of theoretical yield and limiting reactants/reagents.
Derive the theoretical yield for a reaction under specified conditions.
Calculate the percent yield for a reaction.
The relative amounts of reactants and products represented in a balanced chemical equation are often referred to as stoichiometric
amounts. All the exercises of the preceding module involved stoichiometric amounts of reactants. For example, when calculating
the amount of product generated from a given amount of reactant, it was assumed that any other reactants required were available
in stoichiometric amounts (or greater). In this module, more realistic situations are considered, in which reactants are not present in
stoichiometric amounts.
Limiting Reactant
Consider another food analogy, making grilled cheese sandwiches (Figure 5.3.1):
1 slice of cheese + 2 slices of bread → 1 sandwich (5.3.1)
Stoichiometric amounts of sandwich ingredients for this recipe are bread and cheese slices in a 2:1 ratio. Provided with 28 slices of
bread and 11 slices of cheese, one may prepare 11 sandwiches per the provided recipe, using all the provided cheese and having six
slices of bread left over. In this scenario, the number of sandwiches prepared has been limited by the number of cheese slices, and
the bread slices have been provided in excess.
Figure 5.3.1 : Sandwich making can illustrate the concepts of limiting and excess reactants.
Consider this concept now with regard to a chemical process, the reaction of hydrogen with chlorine to yield hydrogen chloride:
Cl (g) → 2 HCl(g) (5.3.2)
2
The balanced equation shows the 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 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 nonreacted.
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
5.3.1 https://chem.libretexts.org/@go/page/217272
2 mol HCl
mol HCl produced = 3 mol H2 × = 6 mol HCl (5.3.3)
1 mol H2
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 5.3.2).
Figure 5.3.2 : 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.
5.3.2 https://chem.libretexts.org/@go/page/217272
Reactants, Products and L
Molecules
Sandwiches
1 mol N2
mol N2 = 1.50 g N2 × = 0.0535 mol N2
28.02 g N2
5.3.3 https://chem.libretexts.org/@go/page/217272
The provided Si:N2 molar ratio is:
0.0712 mol Si 1.33 mol Si
=
0.0535 mol N2 1 mol N2
Comparing these ratios shows that Si is provided in a less-than-stoichiometric amount, and so is the limiting reactant.
Alternatively, compute the amount of product expected for complete reaction of each of the provided reactants. The 0.0712 moles
of silicon would yield
1 mol Si3 N4
mol Si3 N4 produced = 0.0712 mol Si × = 0.0237 mol Si3 N4
3 mol Si
Since silicon yields the lesser amount of product, it is the limiting reactant.
Exercise 5.3.1
Which is the limiting reactant when 5.00 g of H2 and 10.0 g of O2 react and form water?
Answer
O2
Percent Yield
5.3.4 https://chem.libretexts.org/@go/page/217272
Actual and theoretical yields may be expressed as masses or molar amounts (or any other appropriate property; e.g., volume, if the
product is a gas). As long as both yields are expressed using the same units, these units will cancel when percent yield is calculated.
Using this theoretical yield and the provided value for actual yield, the percent yield is calculated to be
actual yield
percent yield = ( ) × 100
theoretical yield
0.392 g Cu
percent yield = ( ) × 100
0.5072 g Cu
= 77.3%
Exercise 5.3.2
What is the percent yield of a reaction that produces 12.5 g of the gas Freon CF2Cl2 from 32.9 g of CCl4 and excess HF?
CCl + 2 HF → CF Cl + 2 HCl
4 2 2
Answer
48.3%
5.3.5 https://chem.libretexts.org/@go/page/217272
products. The atom economy of a process is a measure of this efficiency, defined as the percentage by mass of the final product
of a synthesis relative to the masses of all the reactants used:
mass of product
atom economy = × 100% (5.3.7)
mass of reactants
Though the definition of atom economy at first glance appears very similar to that for percent yield, be aware that this property
represents a difference in the theoretical efficiencies of different chemical processes. The percent yield of a given chemical
process, on the other hand, evaluates the efficiency of a process by comparing the yield of product actually obtained to the
maximum yield predicted by stoichiometry.
The synthesis of the common nonprescription pain medication, ibuprofen, nicely illustrates the success of a green chemistry
approach (Figure 5.3.3). First marketed in the early 1960s, ibuprofen was produced using a six-step synthesis that required 514
g of reactants to generate each mole (206 g) of ibuprofen, an atom economy of 40%. In the 1990s, an alternative process was
developed by the BHC Company (now BASF Corporation) that requires only three steps and has an atom economy of ~80%,
nearly twice that of the original process. The BHC process generates significantly less chemical waste; uses less-hazardous and
recyclable materials; and provides significant cost-savings to the manufacturer (and, subsequently, the consumer). In recognition
of the positive environmental impact of the BHC process, the company received the Environmental Protection Agency’s
Greener Synthetic Pathways Award in 1997.
Figure 5.3.3 : (a) Ibuprofen is a popular nonprescription pain medication commonly sold as 200 mg tablets. (b) The BHC
process for synthesizing ibuprofen requires only three steps and exhibits an impressive atom economy. (credit a: modification of
work by Derrick Coetzee)
Summary
When reactions are carried out using less-than-stoichiometric quantities of reactants, the amount of product generated will be
determined by the limiting reactant. The amount of product generated by a chemical reaction is its actual yield. This yield is often
less than the amount of product predicted by the stoichiometry of the balanced chemical equation representing the reaction (its
theoretical yield). The extent to which a reaction generates the theoretical amount of product is expressed as its percent yield.
Key Equations
actual yield
percent yield = ( ) × 100
theoretical yield
Glossary
actual yield
amount of product formed in a reaction
excess reactant
5.3.6 https://chem.libretexts.org/@go/page/217272
reactant present in an amount greater than required by the reaction stoichiometry
limiting reactant
reactant present in an amount lower than required by the reaction stoichiometry, thus limiting the amount of product generated
percent yield
measure of the efficiency of a reaction, expressed as a percentage of the theoretical yield
theoretical yield
amount of product that may be produced from a given amount of reactant(s) according to the reaction stoichiometry
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Fuse School, Open Educational Resource free of charge, under a Creative Commons License: Attribution-NonCommercial CC
BY-NC (View License Deed: https://creativecommons.org/licenses/by-nc/4.0/)
5.3: Calculating Reaction Yields is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
5.3.7 https://chem.libretexts.org/@go/page/217272
5.3.1: Practice Problems Calculating Reaction Yields
PROBLEM 5.3.1.1
The following quantities are placed in a container: 1.5 × 1024 molecules of diatomic hydrogen, 1.0 mol of sulfur, and 88.0 g of
diatomic oxygen.
a. What is the total mass in grams for the collection of all three elements?
b. What is the total number of moles of atoms for the three elements?
c. If the mixture of the three elements formed a compound with molecules that contain two hydrogen atoms, one sulfur atom,
and four oxygen atoms, which substance is consumed first?
d. How many atoms or molecules of each remaining element would remain unreacted in the change described above ?
Answer a
4.98 g H2, 32 g S; 124.98 g total
Answer b
2.49 mol H2, 2.75 mol O2; 6.24 mol total
Answer c
S is the limiting reactant (what would be consumed first)
Answer d
8.978 × 1023 molecules of H2 and 4.6 × 1023 molecules of O2 remain
Problem 5.3.1
PROBLEM 5.3.1.2
What is the limiting reactant in a reaction that produces sodium chloride from 8 g of sodium and 8 g of diatomic chlorine?
Answer
The limiting reactant is Cl2
PROBLEM 5.3.1.3
Which of the postulates of Dalton's atomic theory explains why we can calculate a theoretical yield for a chemical reaction?
Answer
Postulate 4 (A compound consists of atoms of two or more elements combined in a small, whole-number ratio. In a given
compound, the numbers of atoms of each of its elements are always present in the same ratio).
5.3.1.1 https://chem.libretexts.org/@go/page/217274
PROBLEM 5.3.1.4
A student isolated 25 g of a compound following a procedure that would theoretically yield 81 g. What was his percent yield?
Answer
Percent yield = 31%
PROBLEM 5.3.1.5
A sample of 0.53 g of carbon dioxide was obtained by heating 1.31 g of calcium carbonate. What is the percent yield for this
reaction?
CaCO (s) → CaO(s) + CO (s) (5.3.1.1)
3 2
Answer
Percent yield = 91.9%
Problem 5.3.5
PROBLEM 5.3.1.6
Freon-12, CCl2F2, is prepared from CCl4 by reaction with HF. The other product of this reaction is HCl. Outline the steps
needed to determine the percent yield of a reaction that produces 12.5 g of CCl2F2 from 32.9 g of CCl4. Freon-12 has been
banned and is no longer used as a refrigerant because it catalyzes the decomposition of ozone and has a very long lifetime in the
atmosphere. Determine the percent yield.
Answer
g CCl → mol CCl → mol CCl F → g CCl F
4 4 2 2 2 2
PROBLEM 5.3.1.7
Citric acid, C6H8O7, a component of jams, jellies, and fruity soft drinks, is prepared industrially via fermentation of sucrose by
the mold Aspergillus niger. The equation representing this reaction is
C H O +H O+3 O → 2C H O +4 H O (5.3.1.2)
12 22 11 2 2 6 8 7 2
What mass of citric acid is produced from exactly 1 metric ton (1.000 × 103 kg) of sucrose (C12H22O11) if the yield is 92.30%?
Answer
1036 kg citric acid
5.3.1.2 https://chem.libretexts.org/@go/page/217274
Click here to see a video of the solution
Problem 5.3.7
PROBLEM 5.3.1.8
Toluene, C6H5CH3, is oxidized by air under carefully controlled conditions to benzoic acid, C6H5CO2H, which is used to
prepare the food preservative sodium benzoate, C6H5CO2Na. What is the percent yield of a reaction that converts 1.000 kg of
toluene to 1.21 kg of benzoic acid?
2 C H CH +3 O → 2 C H CO H + 2 H O (5.3.1.3)
6 5 3 2 6 5 2 2
Answer
percent yield = 91.3%
PROBLEM 5.3.1.9
In a laboratory experiment, the reaction of 3.0 mol of H2 with 2.0 mol of I2 produced 1.0 mol of HI. Determine the theoretical
yield in grams and the percent yield for this reaction.
Answer
512 g (theoretical yield); percent yield = 25%
Problem 5.3.9
PROBLEM 5.3.1.10
What is the limiting reactant when 1.50 g of lithium and 1.50 g of nitrogen combine to form lithium nitride, a component of
advanced batteries, according to the following unbalanced equation?
Li + N → Li N (5.3.1.4)
2 3
5.3.1.3 https://chem.libretexts.org/@go/page/217274
Answer
6 Li + N → 2 Li N (5.3.1.5)
2 3
1 mole Li 2 mole Li N
3
1.50g Li × × = 0.0720 moles Li N (5.3.1.6)
3
6.94g Li 6 mole Li
1 mole N 2 mole Li N
2 3
1.50g N × × = 0.107 moles Li N (5.3.1.7)
2 3
28.02g N 1 mole N
2 2
PROBLEM 5.3.1.11
Uranium can be isolated from its ores by dissolving it as UO2(NO3)2, then separating it as solid UO2(C2O4)·3H2O. Addition of
0.4031 g of sodium oxalate, Na2C2O4, to a solution containing 1.481 g of uranyl nitrate, UO2(NO3)2, yields 1.073 g of solid
UO2(C2O4)·3H2O.
Na C O + UO (NO ) + 3 H O ⟶ UO (C O ) ⋅ 3 H O + 2 NaNO (5.3.1.8)
2 2 4 2 3 2 2 2 2 4 2 3
Determine the limiting reactant and the percent yield of this reaction.
Answer
Na2C2O4 is the limiting reactant. percent yield = 86.6%
Problem 5.3.11
PROBLEM 5.3.1.12
How many molecules of C2H4Cl2 can be prepared from 15 C2H4 molecules and 8 Cl2 molecules?
Answer
Only 8 molecules can be formed
PROBLEM 5.3.1.13
How many molecules of the sweetener saccharin can be prepared from 30 C atoms, 25 H atoms, 12 O atoms, 8 S atoms, and 14
N atoms?
5.3.1.4 https://chem.libretexts.org/@go/page/217274
Answer
Only four molecules can be made.
PROBLEM 5.3.1.14
The phosphorus pentoxide used to produce phosphoric acid for cola soft drinks is prepared by burning phosphorus in oxygen.
a. What is the limiting reactant when 0.200 mol of P4 and 0.200 mol of O2 react according to
P +5 O → P O (5.3.1.9)
4 2 4 10
b. Calculate the percent yield if 10.0 g of P4O10 is isolated from the reaction.
Answer a
O2 is the limiting reactant
Answer b
percent yield = 88%
PROBLEM 5.3.1.15
Would you agree to buy 1 trillion (1,000,000,000,000) gold atoms for $5? Explain why or why not. Find the current price of
gold at http://money.cnn.com/data/commodities/ (1 troy ounce = 31.1 g)
Answer
This amount cannot be weighted by ordinary balances and is worthless.
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
5.3.1: Practice Problems Calculating Reaction Yields is shared under a not declared license and was authored, remixed, and/or curated by
LibreTexts.
5.3.1.5 https://chem.libretexts.org/@go/page/217274
CHAPTER OVERVIEW
6: Aqueous Reactions
6.1: Solutions and Solution Concentration
6.1.1: Practice Problems- Solution Concentration
6.2: Solutions Chemistry
6.2.1: Practice Problems- Solutions Chemistry
6.3: Precipitation Reactions
6.3.1: Practice Problems- Precipitation
6.4: Classifying Chemical Reactions (Acids and Bases)
6.4.1: Classifying Chemical Reactions (Acids and Bases) (Problems)
6.5: Classifying Chemical Reactions (Redox)
6.5.1: Classifying Chemical Reactions (Redox) (Problems)
6: Aqueous Reactions is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
1
6.1: Solutions and Solution Concentration
Skills to Develop
Describe the fundamental properties of solutions
Calculate solution concentrations using molarity and mass-mass-percent
Perform dilution calculations using the dilution equation
In preceding sections, we focused on the composition of substances: samples of matter that contain only one type of element or
compound. However, mixtures—samples of matter containing two or more substances physically combined—are more commonly
encountered in nature than are pure substances. Similar to a pure substance, the relative composition of a mixture plays an
important role in determining its properties. The relative amount of oxygen in a planet’s atmosphere determines its ability to sustain
aerobic life. The relative amounts of iron, carbon, nickel, and other elements in steel (a mixture known as an “alloy”) determine its
physical strength and resistance to corrosion. The relative amount of the active ingredient in a medicine determines its effectiveness
in achieving the desired pharmacological effect. The relative amount of sugar in a beverage determines its sweetness (Figure
6.1.1). In this section, we will describe one of the most common ways in which the relative compositions of mixtures may be
quantified.
Figure 6.1.1 : Sugar is one of many components in the complex mixture known as coffee. The amount of sugar in a given amount
of coffee is an important determinant of the beverage’s sweetness. (credit: Jane Whitney)
Solutions
6.1.1 https://chem.libretexts.org/@go/page/217277
The relative amount of a given solution component is known as its concentration. Often, though not always, a solution contains one
component with a concentration that is significantly greater than that of all other components. This component is called the solvent
and may be viewed as the medium in which the other components are dispersed, or dissolved. Solutions in which water is the
solvent are, of course, very common on our planet. A solution in which water is the solvent is called an aqueous solution.
A solute is a component of a solution that is typically present at a much lower concentration than the solvent. Solute concentrations
are often described with qualitative terms such as dilute (of relatively low concentration) and concentrated (of relatively high
concentration).
Concentrations of Solutions
Concentrations may be quantitatively assessed using a wide variety of measurement units, each convenient for particular
applications. Some of the common concetrations measurements seen in the science labs include, parts-per-million (ppm), parts-per-
billion (ppb), mass-mass-percent (m/m %), molarity (M), molality (m) and mole percent. In this unit we will use two of the more
common units of measure, Molarity, and mass-mass percent.
Molarity
Molarity (M) is a useful concentration unit for many applications in chemistry. Molarity is defined as the number of moles of solute
in exactly 1 liter (1 L) of the solution:
mol solute
M = (6.1.1)
L solution
The units of molarity are therefore moles per liter of solution (mol/L), abbreviated as M . An aqueous solution that contains 1 mol
(342 g) of sucrose in enough water to give a final volume of 1.00 L has a sucrose concentration of 1.00 mol/L or 1.00 M. In
chemical notation, square brackets around the name or formula of the solute represent the molar concentration of a solute.
Therefore,
[sucrose] = 1.00 M (6.1.2)
6.1.2 https://chem.libretexts.org/@go/page/217277
mol solute
M =
L solution
0.133 mol
=
1 L
355 mL ×
1000 mL
= 0.375 M
Exercise 6.1.1
A teaspoon of table sugar contains about 0.01 mol sucrose. What is the molarity of sucrose if a teaspoon of sugar has been
dissolved in a cup of tea with a volume of 200 mL?
Answer
0.05 M
mol sugar 1 L
mol solute = 0.375 × (10 mL × ) = 0.004 mol sugar
L 1000 mL
Exercise 6.1.2
What volume (mL) of the sweetened tea described in Example 6.1.1 contains the same amount of sugar (mol) as 10 mL of the
soft drink in this example?
Answer
80 mL
25.2 g of acetic acid. What is the concentration of the acetic acid solution in units of molarity?
6.1.3 https://chem.libretexts.org/@go/page/217277
Solution
As in previous examples, the definition of molarity is the primary equation used to calculate the quantity sought. In this case, the
mass of solute is provided instead of its molar amount, so we must use the solute’s molar mass to obtain the amount of solute in
moles:
1 mol CH CO H
3 2
25.2 g CH CO H ×
3 2
mol solute 60.052 g CH CO H
3 2
M = = = 0.839 M (6.1.3)
L solution 0.500 L solution
mol solute
M = = 0.839 M (6.1.4)
L solution
Exercise 6.1.3
Calculate the molarity of 6.52 g of C oC l (128.9 g/mol) dissolved in an aqueous solution with a total volume of 75.0 mL.
2
Answer
0.674 M
mol NaCl
mol solute = 5.30 × 0.250 L = 1.325 mol NaCl (6.1.8)
L
Exercise 6.1.4
How many grams of C aC l (110.98 g/mol) are contained in 250.0 mL of a 0.200-M solution of calcium chloride?
2
Answer
5.55 g C aC l2
When performing calculations stepwise, as in Example 6.1.3, it is important to refrain from rounding any intermediate calculation
results, which can lead to rounding errors in the final result. In Example 6.1.4, the molar amount of NaCl computed in the first
step, 1.325 mol, would be properly rounded to 1.32 mol if it were to be reported; however, although the last digit (5) is not
significant, it must be retained as a guard digit in the intermediate calculation. If we had not retained this guard digit, the final
calculation for the mass of NaCl would have been 77.1 g, a difference of 0.3 g.
In addition to retaining a guard digit for intermediate calculations, we can also avoid rounding errors by performing computations
in a single step (Example 6.1.5). This eliminates intermediate steps so that only the final result is rounded.
6.1.4 https://chem.libretexts.org/@go/page/217277
Example 6.1.5 : Determining the Volume of Solution
In Example , we found the typical concentration of vinegar to be 0.839 M. What volume of vinegar contains 75.6 g of
6.1.3
acetic acid?
Solution
First, use the molar mass to calculate moles of acetic acid from the given mass:
mol solute
g solute × = mol solute (6.1.10)
g solute
Then, use the molarity of the solution to calculate the volume of solution containing this molar amount of solute:
L solution
mol solute × = L solution (6.1.11)
mol solute
mol CH CO H L solution
3 2
75.6 g CH CO H ( )( ) = 1.50 L solution (6.1.13)
3 2
60.05 g 0.839 mol CH CO H
3 2
Exercise 6.1.5 :
What volume of a 1.50-M KBr solution contains 66.0 g KBr?
Answer
0.370 L
Mass-mass percent
Percentages are also commonly used to express the composition of mixtures, including solutions. The mass percentage (m/m%)of
a solution component is defined as the ratio of the component’s mass to the solution’s mass, expressed as a percentage:
mass of solute
% m/m = × 100% (6.1.14)
mass of solution
Mass percentage is also referred to by similar names such as percent mass, percent weight, weight/weight percent, and other
variations on this theme. The most common symbol for mass percentage is simply the percent sign, %, although more detailed
symbols are often used including %mass, %weight, and (w/w)%. Use of these more detailed symbols can prevent confusion of
mass percentages with other types of percentages, such as volume percentages (to be discussed later in this section).
Mass percentages are popular concentration units for consumer products. The label of a typical liquid bleach bottle (Figure 3.17)
cites the concentration of its active ingredient, sodium hypochlorite (NaOCl), as being 7.4%. A 100.0-g sample of bleach would
6.1.5 https://chem.libretexts.org/@go/page/217277
therefore contain 7.4 g of NaOCl.
The computed mass percentage agrees with our rough estimate (it’s a bit less than 0.1%).
Note that while any mass unit may be used to compute a mass percentage (mg, g, kg, oz, and so on), the same unit must be used
for both the solute and the solution so that the mass units cancel, yielding a dimensionless ratio. In this case, we converted the
units of solute in the numerator from mg to g to match the units in the denominator. We could just as easily have converted the
denominator from g to mg instead. As long as identical mass units are used for both solute and solution, the computed mass
percentage will be correct.
Exercise 6.1.6 :
A bottle of a tile cleanser contains 135 g of HCl and 775 g of water. What is the percent by mass of HCl in this cleanser?
Answer
14.8%
6.1.6 https://chem.libretexts.org/@go/page/217277
“Concentrated” hydrochloric acid is an aqueous solution of 37.2% HCl that is commonly used as a laboratory reagent. The
density of this solution is 1.19 g/mL. What mass of HCl is contained in 0.500 L of this solution?
Solution
The HCl concentration is near 40%, so a 100-g portion of this solution would contain about 40 g of HCl. Since the solution
density isn’t greatly different from that of water (1 g/mL), a reasonable estimate of the HCl mass in 500 g (0.5 L) of the solution
is about five times greater than that in a 100 g portion, or 5 × 40 = 200 g . To derive the mass of solute in a solution from its
mass percentage, we need to know the corresponding mass of the solution. Using the solution density given, we can convert the
solution’s volume to mass, and then use the given mass percentage to calculate the solute mass. This mathematical approach is
outlined in this flowchart:
For proper unit cancellation, the 0.500-L volume is converted into 500 mL, and the mass percentage is expressed as a ratio, 37.2
g HCl/g solution:
1.19 g solution 37.2 g HCl
500 mL solution ( )( ) = 221 g HCl (6.1.17)
mL solution 100 g solution
This mass of HCl is consistent with our rough estimate of approximately 200 g.
Exercise 6.1.7
What volume of concentrated HCl solution contains 125 g of HCl?
Answer
282 mL
Dilution of Solutions
6.1.7 https://chem.libretexts.org/@go/page/217277
Figure 6.1.5 : Both solutions contain the same mass of copper nitrate. The solution on the right is more dilute because the copper
nitrate is dissolved in more solvent. (credit: Mark Ott).
Dilution is also a common means of preparing solutions of a desired concentration. By adding solvent to a measured portion of a
more concentrated stock solution, we can achieve a particular concentration. For example, commercial pesticides are typically sold
as solutions in which the active ingredients are far more concentrated than is appropriate for their application. Before they can be
used on crops, the pesticides must be diluted. This is also a very common practice for the preparation of a number of common
laboratory reagents (Figure 6.1.3).
Figure 6.1.6 : A solution of KM nO is prepared by mixing water with 4.74 g of KMnO4 in a flask. (credit: modification of work
4
by Mark Ott)
A simple mathematical relationship can be used to relate the volumes and concentrations of a solution before and after the dilution
process. According to the definition of molarity, the molar amount of solute in a solution is equal to the product of the solution’s
molarity and its volume in liters:
n = ML (6.1.18)
Expressions like these may be written for a solution before and after it is diluted:
n1 = M1 L1 (6.1.19)
n2 = M2 L2 (6.1.20)
where the subscripts “1” and “2” refer to the solution before and after the dilution, respectively. Since the dilution process does not
change the amount of solute in the solution,n1 = n2. Thus, these two equations may be set equal to one another:
M1 L1 = M2 L2 (6.1.21)
This relation is commonly referred to as the dilution equation. Although we derived this equation using molarity as the unit of
concentration and liters as the unit of volume, other units of concentration and volume may be used, so long as the units properly
cancel per the factor-label method. Reflecting this versatility, the dilution equation is often written in the more general form:
C1 V1 = C2 V2 (6.1.22)
6.1.8 https://chem.libretexts.org/@go/page/217277
Phet Simulation
Solute Solution
Amount Volume
(moles) (Liters)
lots full
Drink mix
low
none
Molarity
C1 V1
C2 = (6.1.24)
V2
Since the stock solution is being diluted by more than two-fold (volume is increased from 0.85 L to 1.80 L), we would expect
the diluted solution’s concentration to be less than one-half 5 M. We will compare this ballpark estimate to the calculated result
6.1.9 https://chem.libretexts.org/@go/page/217277
to check for any gross errors in computation (for example, such as an improper substitution of the given quantities). Substituting
the given values for the terms on the right side of this equation yields:
mol
0.850 L × 5.00
L
C2 = = 2.36 M (6.1.25)
1.80 L
This result compares well to our ballpark estimate (it’s a bit less than one-half the stock concentration, 5 M).
Exercise 6.1.8
What is the concentration of the solution that results from diluting 25.0 mL of a 2.04-M solution of CH3OH to 500.0 mL?
Answer
0.102 M C H 3 OH
C1 V1
V2 = (6.1.27)
C2
Since the diluted concentration (0.12 M) is slightly more than one-fourth the original concentration (0.45 M), we would expect
the volume of the diluted solution to be roughly four times the original volume, or around 44 mL. Substituting the given values
and solving for the unknown volume yields:
(0.45 M )(0.011 L)
V2 = (6.1.28)
(0.12 M )
V2 = 0.041 L (6.1.29)
The volume of the 0.12-M solution is 0.041 L (41 mL). The result is reasonable and compares well with our rough estimate.
Exercise 6.1.9
A laboratory experiment calls for 0.125 M H N O . What volume of 0.125 M H N O can be prepared from 0.250 L of 1.88 M
3 3
HN O ?3
Answer
3.76 L
6.1.10 https://chem.libretexts.org/@go/page/217277
C2 V2
V1 = (6.1.31)
C1
Since the concentration of the diluted solution 0.100 M is roughly one-sixteenth that of the stock solution (1.59 M), we would
expect the volume of the stock solution to be about one-sixteenth that of the diluted solution, or around 0.3 liters. Substituting
the given values and solving for the unknown volume yields:
(0.100 M )(5.00 L)
V1 = (6.1.32)
1.59 M
V1 = 0.314 L (6.1.33)
Thus, we would need 0.314 L of the 1.59-M solution to prepare the desired solution. This result is consistent with our rough
estimate.
Exercise 6.1.10
What volume of a 0.575-M solution of glucose, C6H12O6, can be prepared from 50.00 mL of a 3.00-M glucose solution?
Answer
0.261 L
Summary
Solutions are homogeneous mixtures. Many solutions contain one component, called the solvent, in which other components,
called solutes, are dissolved. An aqueous solution is one for which the solvent is water. The concentration of a solution is a measure
of the relative amount of solute in a given amount of solution. Concentrations may be measured using various units, with one very
useful unit being molarity, defined as the number of moles of solute per liter of solution. The solute concentration of a solution may
be decreased by adding solvent, a process referred to as dilution. The dilution equation is a simple relation between concentrations
and volumes of a solution before and after dilution.
6.1.11 https://chem.libretexts.org/@go/page/217277
What Are Dilutions | Chemical Calculatio…
Calculatio…
Video 6.1.5 : An review of calculating dilutions. Also, orange squash is a drink in England.
Key Equations
mol solute
M =
L solution
C1V1 = C2V2
Glossary
aqueous solution
solution for which water is the solvent
concentrated
qualitative term for a solution containing solute at a relatively high concentration
concentration
quantitative measure of the relative amounts of solute and solvent present in a solution
dilute
qualitative term for a solution containing solute at a relatively low concentration
dilution
process of adding solvent to a solution in order to lower the concentration of solutes
dissolved
describes the process by which solute components are dispersed in a solvent
molarity (M)
unit of concentration, defined as the number of moles of solute dissolved in 1 liter of solution
solute
solution component present in a concentration less than that of the solvent
solvent
solution component present in a concentration that is higher relative to other components
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
6.1.12 https://chem.libretexts.org/@go/page/217277
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Fuse School, Open Educational Resource free of charge, under a Creative Commons License: Attribution-NonCommercial CC
BY-NC (View License Deed: https://creativecommons.org/licenses/by-nc/4.0/)
Modified by Joshua Halpern (Howard University)
6.1: Solutions and Solution Concentration is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
6.1.13 https://chem.libretexts.org/@go/page/217277
6.1.1: Practice Problems- Solution Concentration
PROBLEM 6.1.1.1
Explain what changes and what stays the same when 1.00 L of a solution of NaCl is diluted to 1.80 L.
Answer
The number of moles always stays the same in a dilution.
The concentration and the volumes change in a dilution.
PROBLEM 6.1.1.2
What does it mean when we say that a 200-mL sample and a 400-mL sample of a solution of salt have the same molarity? In
what ways are the two samples identical? In what ways are these two samples different?
Answer
The two samples contain the same proportion of moles of salt to liters of solution, but have different numbers of actual
moles.
PROBLEM 6.1.1.3
Determine the molarity for each of the following solutions:
a. 0.444 mol of CoCl2 in 0.654 L of solution
b. 98.0 g of phosphoric acid, H3PO4, in 1.00 L of solution
c. 0.2074 g of calcium hydroxide, Ca(OH)2, in 40.00 mL of solution
d. 10.5 kg of Na2SO4·10H2O in 18.60 L of solution
e. 7.0 × 10−3 mol of I2 in 100.0 mL of solution
f. 1.8 × 104 mg of HCl in 0.075 L of solution
Answer a
0.679 M
Answer b
1.00 M
Answer c
0.06998 M
Answer d
1.75 M
Answer e
0.070 M
Answer f
6.6 M
6.1.1.1 https://chem.libretexts.org/@go/page/217282
Problem 6.1.3
PROBLEM 6.1.1.4
Determine the molarity of each of the following solutions:
a. 1.457 mol KCl in 1.500 L of solution
b. 0.515 g of H2SO4 in 1.00 L of solution
c. 20.54 g of Al(NO3)3 in 1575 mL of solution
d. 2.76 kg of CuSO4·5H2O in 1.45 L of solution
e. 0.005653 mol of Br2 in 10.00 mL of solution
f. 0.000889 g of glycine, C2H5NO2, in 1.05 mL of solution
Answer a
0.9713 M
Answer b
5.25 × 10-3 M
Answer c
6.122 × 10-2 M
Answer d
7.62 M
Answer e
0.5653 M
Answer f
1.13 × 10-2 M
PROBLEM 6.1.1.5
Calculate the number of moles and the mass of the solute in each of the following solutions:
6.1.1.2 https://chem.libretexts.org/@go/page/217282
(a) 2.00 L of 18.5 M H2SO4, concentrated sulfuric acid
(b) 100.0 mL of 3.8 × 10−5 M NaCN, the minimum lethal concentration of sodium cyanide in blood serum
(c) 5.50 L of 13.3 M H2CO, the formaldehyde used to “fix” tissue samples
(d) 325 mL of 1.8 × 10−6 M FeSO4, the minimum concentration of iron sulfate detectable by taste in drinking water
Answer a
37.0 mol H2SO4
3.63 × 103 g H2SO4
Answer b
3.8 × 10−6 mol NaCN
1.9 × 10−4 g NaCN
Answer c
73.2 mol H2CO
2.20 kg H2CO
Answer d
5.9 × 10−7 mol FeSO4
8.9 × 10−5 g FeSO4
Problem 6.1.5
PROBLEM 6.1.1.6
Calculate the molarity of each of the following solutions:
(a) 0.195 g of cholesterol, C27H46O, in 0.100 L of serum, the average concentration of cholesterol in human serum
(b) 4.25 g of NH3 in 0.500 L of solution, the concentration of NH3 in household ammonia
(c) 1.49 kg of isopropyl alcohol, C3H7OH, in 2.50 L of solution, the concentration of isopropyl alcohol in rubbing alcohol
(d) 0.029 g of I2 in 0.100 L of solution, the solubility of I2 in water at 20 °C
Answer a
6.1.1.3 https://chem.libretexts.org/@go/page/217282
5.04 × 10−3 M
Answer b
0.499 M
Answer c
9.92 M
Answer d
1.1 × 10−3 M
PROBLEM 6.1.1.7
There is about 1.0 g of calcium, as Ca2+, in 1.0 L of milk. What is the molarity of Ca2+ in milk?
Answer
0.025 M
Problem 6.1.7
PROBLEM 6.1.1.8
What volume of a 1.00-M Fe(NO3)3 solution can be diluted to prepare 1.00 L of a solution with a concentration of 0.250 M?
Answer
0.250 L
PROBLEM 6.1.1.9
If 0.1718 L of a 0.3556-M C3H7OH solution is diluted to a concentration of 0.1222 M, what is the volume of the resulting
solution?
Answer
0.5000 L
6.1.1.4 https://chem.libretexts.org/@go/page/217282
Problem 6.1.9
PROBLEM 6.1.1.10
What volume of a 0.33-M C12H22O11 solution can be diluted to prepare 25 mL of a solution with a concentration of 0.025 M?
Answer
1.9 mL
PROBLEM 6.1.1.11
What is the concentration of the NaCl solution that results when 0.150 L of a 0.556-M solution is allowed to evaporate until the
volume is reduced to 0.105 L?
Answer
0.794 M
Problem 6.1.11
PROBLEM 6.1.1.12
What is the molarity of the diluted solution when each of the following solutions is diluted to the given final volume?
a. 1.00 L of a 0.250-M solution of Fe(NO3)3 is diluted to a final volume of 2.00 L
b. 0.5000 L of a 0.1222-M solution of C3H7OH is diluted to a final volume of 1.250 L
c. 2.35 L of a 0.350-M solution of H3PO4 is diluted to a final volume of 4.00 L
d. 22.50 mL of a 0.025-M solution of C12H22O11 is diluted to 100.0 mL
6.1.1.5 https://chem.libretexts.org/@go/page/217282
Answer a
0.125 M
Answer b
0.04888 M
Answer c
0.206 M
Answer d
0.0056 M
PROBLEM 6.1.1.13
What is the final concentration of the solution produced when 225.5 mL of a 0.09988-M solution of Na2CO3 is allowed to
evaporate until the solution volume is reduced to 45.00 mL?
Answer
0.5005 M
Problem 6.1.13
PROBLEM 6.1.1.14
A 2.00-L bottle of a solution of concentrated HCl was purchased for the general chemistry laboratory. The solution contained
868.8 g of HCl. What is the molarity of the solution?
Answer
11.9 M
PROBLEM 6.1.1.15
An experiment in a general chemistry laboratory calls for a 2.00-M solution of HCl. How many mL of 11.9 M HCl would be
required to make 250 mL of 2.00 M HCl?
Answer
42.0 mL
6.1.1.6 https://chem.libretexts.org/@go/page/217282
Click here to see a video of the solution
Problem 6.1.15
PROBLEM 6.1.1.16
What volume of a 0.20-M K2SO4 solution contains 57 g of K2SO4?
Answer
1.6 L
PROBLEM 6.1.1.17
The US Environmental Protection Agency (EPA) places limits on the quantities of toxic substances that may be discharged into
the sewer system. Limits have been established for a variety of substances, including hexavalent chromium, which is limited to
0.50 mg/L. If an industry is discharging hexavalent chromium as potassium dichromate (K2Cr2O7), what is the maximum
permissible molarity of that substance?
Answer
4.8 × 10−6 M
Problem 6.1.17
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
6.1.1.7 https://chem.libretexts.org/@go/page/217282
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
6.1.1: Practice Problems- Solution Concentration is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
6.1.1.8 https://chem.libretexts.org/@go/page/217282
6.2: Solutions Chemistry
Skills to Develop
Describe the basic properties of solutions and how they form
Define and give examples of electrolytes
Predict the solubility of common inorganic compounds by using solubility rules
Write and balance chemical equations in molecular, total ionic, and net ionic formats.
The previous unit introduced solutions, defined as homogeneous mixtures of two or more substances. Often, one component of a
solution is present at a significantly greater concentration, in which case it is called the solvent. The other components of the
solution present in relatively lesser concentrations are called solutes. Sugar is a covalent solid composed of sucrose molecules,
C12 H 22O . When this compound dissolves in water, its molecules become uniformly distributed among the molecules of water:
11
The subscript “aq” in the equation signifies that the sucrose molecules are solutes and are therefore individually dispersed
throughout the aqueous solution (water is the solvent). Although sucrose molecules are heavier than water molecules, they remain
dispersed throughout the solution; gravity does not cause them to “settle out” over time.
Potassium dichromate, K C r O , is an ionic compound composed of colorless potassium ions, K , and orange dichromate ions,
2 2 7
+
Cr O
2 7
. When a small amount of solid potassium dichromate is added to water, the compound dissolves and dissociates to yield
2−
potassium ions and dichromate ions uniformly distributed throughout the mixture (Figure 6.2.1), as indicated in this equation:
+ 2−
K2 C r2 O7(s) ⟶ 2K (aq)
+ Cr2 O (6.2.2)
7 (aq)
As with the mixture of sugar and water, this mixture is also an aqueous solution. Its solutes, potassium and dichromate ions, remain
individually dispersed among the solvent (water) molecules.
Figure 6.2.1 : When potassium dichromate (K C r O ) is mixed with water, it forms a homogeneous orange solution. (credit:
2 2 7
Click to Run
Water is used so often as a solvent that the word solution has come to imply an aqueous solution to many people. However, almost
any gas, liquid, or solid can act as a solvent. Many alloys are solid solutions of one metal dissolved in another; for example, US
five-cent coins contain nickel dissolved in copper. Air is a gaseous solution, a homogeneous mixture of nitrogen, oxygen, and
6.2.1 https://chem.libretexts.org/@go/page/217278
several other gases. Oxygen (a gas), alcohol (a liquid), and sugar (a solid) all dissolve in water (a liquid) to form liquid solutions.
Table 6.2.1 gives examples of several different solutions and the phases of the solutes and solvents.
Table 6.2.1: Different Types of Solutions
Solution Solute Solvent
6.2.2 https://chem.libretexts.org/@go/page/217278
Figure 6.2.2 : 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 6.2.2.
Figure 6.2.3 : 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. 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.
6.2.3 https://chem.libretexts.org/@go/page/217278
In other cases, the electrostatic attractions between the ions in a crystal are so large 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.
+ + −
H O(l) H O(l) ⇌ H O (aq) + OH (aq) (6.2.3)
2 2 3
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.
Solubility
Solubility is the tendency of one substance to dissolve in another. In chemistry, the term solubility is often used to denote a
substance's aqueous solubility, its ability to dissolve in water.In general if a substance does not dissolve in water it is labeled
insoluble and if it does dissolve in water it is called as soluble. In fact, solubility is much more nuanced than the general terms
soluble and insoluble describe. These nuances will be explored later in this course, for now, we will describe substances as soluble
or insoluble.
A substance's solubility is determined by several factors, including the types and relative strengths of intermolecular attractive
forces that may exist between the substances’ atoms, ions, or molecules. For example, oil and water do not mix because the
intermolecular forces that attract oil particles to one another and the intermolecular forces that attract water molecules together are
much stronger than the intermolecular forces that attract oil particles to water molecules. Intermolecular forces is a subject that will
be expanded on in greater detail later in this course.
6.2.4 https://chem.libretexts.org/@go/page/217278
Why don't oil and water mix? - John Poll…
Poll…
group 1 metal cations (Li+, Na+, K+, Rb+, and Cs+) and ammonium
ion (NH )
+
the halide ions (Cl−, Br−, and I−) halides of Ag+, Hg , and Pb2+
2+
This balanced equation, derived in the usual fashion, is called a molecular equation because it doesn’t explicitly represent the ionic
species that are present in solution. When ionic compounds dissolve in water, they may dissociate into their constituent ions, which
are subsequently dispersed homogenously throughout the resulting solution (a thorough discussion of this important process is
provided in the chapter on solutions). Ionic compounds dissolved in water are, therefore, more realistically represented as
dissociated ions, in this case:
2+ −
CaCl (aq) → Ca (aq) + 2 Cl (aq) (6.2.5)
2
+ −
2 AgNO (aq) → 2 Ag (aq) + 2 NO (aq) (6.2.6)
3 3
6.2.5 https://chem.libretexts.org/@go/page/217278
2+ −
Ca (NO ) (aq) → Ca (aq) + 2 NO3 (aq) (6.2.7)
3 2
Unlike these three ionic compounds, AgCl does not dissolve in water to a significant extent, as signified by its physical state
notation, (s).
Explicitly representing all dissolved ions results in a complete ionic equation. In this particular case, the formulas for the dissolved
ionic compounds are replaced by formulas for their dissociated ions:
2+ − + − 2+ −
Ca (aq) + 2 Cl (aq) + 2 Ag (aq) + 2 NO (aq) → Ca (aq) + 2 NO (aq) + 2 Ag Cl(s) (6.2.8)
3 3
Examining this equation shows that two chemical species are present in identical form on both sides of the arrow, Ca (aq) and 2 +
NO (aq) . These spectator ions—ions whose presence is required to maintain charge neutrality—are neither chemically nor
−
physically changed by the process, and so they may be eliminated from the equation to yield a more succinct representation called a
net ionic equation:
2+ − + − 2+ −
Ca (aq) + 2 Cl (aq) + 2 Ag (aq) + 2 NO (aq) → Ca (aq) + 2 NO (aq) + 2 AgCl(s) (6.2.9)
3 3
− +
2 Cl (aq) + 2 Ag (aq) → 2 AgCl(s) (6.2.10)
Following the convention of using the smallest possible integers as coefficients, this equation is then written:
− +
Cl (aq) + Ag (aq) → AgCl(s) (6.2.11)
This net ionic equation indicates that solid silver chloride may be produced from dissolved chloride and silver(I) ions, regardless of
the source of these ions. These molecular and complete ionic equations provide additional information, namely, the ionic
compounds used as sources of Cl and Ag .
− +
Balance is achieved easily in this case by changing the coefficient for NaOH to 2, resulting in the molecular equation for this
reaction:
+
CO (aq) NaOH(aq) → Na CO (aq) + H O(l)
2 2 2 3 2
The two dissolved ionic compounds, NaOH and Na2CO3, can be represented as dissociated ions to yield the complete ionic
equation:
+ − + 2 −
CO (aq) + 2 Na (aq) + 2 OH (aq) → 2 Na (aq) + CO3 (aq) + H O(l)
2 2
Finally, identify the spectator ion(s), in this case Na+(aq), and remove it from each side of the equation to generate the net ionic
equation:
+ − + 2−
CO (aq) + 2 Na (aq) + 2 OH (aq) → 2 Na (aq) + CO3 (aq) + H O(l)
2 2
− 2−
CO (aq) + 2 OH (aq) → CO3 (aq) + H O(l)
2 2
Exercise 6.2.1
Diatomic chlorine and sodium hydroxide (lye) are commodity chemicals produced in large quantities, along with diatomic
hydrogen, via the electrolysis of brine, according to the following unbalanced equation:
electricity
Write balanced molecular, complete ionic, and net ionic equations for this process.
6.2.6 https://chem.libretexts.org/@go/page/217278
Answer
Balanced molecular equation:
Summary
Glossary
alloy
solid mixture of a metallic element and one or more additional elements
dissociation
physical process accompanying the dissolution of an ionic compound in which the compound’s constituent ions are solvated
and dispersed throughout the solution
6.2.7 https://chem.libretexts.org/@go/page/217278
chemical equation in which all dissolved ionic reactants and products, including spectator ions, are explicitly represented by
formulas for their dissociated ions
electrolyte
substance that produces ions when dissolved in water
insoluble
of relatively low solubility; dissolving only to a slight extent
molecular equation
chemical equation in which all reactants and products are represented as neutral substances
net ionic equation
chemical equation in which only those dissolved ionic reactants and products that undergo a chemical or physical change are
represented (excludes spectator ions)
nonelectrolyte
soluble
of relatively high solubility; dissolving to a relatively large extent
solubility
the extent to which a substance may be dissolved in water, or any solvent
strong electrolyte
substance that dissociates or ionizes completely when dissolved in water
supersaturated
of concentration that exceeds solubility; a nonequilibrium state
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-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
TED-Ed’s commitment to creating lessons worth sharing is an extension of TED’s mission of spreading great ideas. Within
TED-Ed’s growing library of TED-Ed animations, you will find carefully curated educational videos, many of which represent
collaborations between talented educators and animators nominated through the TED-Ed website.
6.2: Solutions Chemistry is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
6.2.8 https://chem.libretexts.org/@go/page/217278
6.2.1: Practice Problems- Solutions Chemistry
PROBLEM 6.2.1.1
How do solutions differ from compounds? From other mixtures?
Answer
A solution can vary in composition, while a compound cannot vary in composition. Solutions are homogeneous at the
molecular level, while other mixtures are heterogeneous.
PROBLEM 6.2.1.2
Give an example of each of the following types of solutions:
a. a gas in a liquid
b. a gas in a gas
c. a solid in a solid
Answer a
CO2 in water
Answer b
O2 in N2 (air)
Answer c
bronze (solution of tin or other metals in copper)
PROBLEM 6.2.1.3
Solutions of hydrogen in palladium may be formed by exposing Pd metal to H2 gas. The concentration of hydrogen in the
palladium depends on the pressure of H2 gas applied, but in a more complex fashion than can be described by Henry’s law.
Under certain conditions, 0.94 g of hydrogen gas is dissolved in 215 g of palladium metal. Determine the molarity of this
solution (solution density = 1.8 g/cm3).
Answer
3.91 M
Problem 6.2.3
PROBLEM 6.2.1.4
6.2.1.1 https://chem.libretexts.org/@go/page/217283
Explain why solutions of HBr in benzene (a nonpolar solvent) are nonconductive, while solutions in water (a polar solvent) are
conductive.
Answer
HBr is an acid and so its molecules react with water molecules to form H3O+ and Br− ions that provide conductivity to the
solution. Though HBr is soluble in benzene, it does not react chemically but remains dissolved as neutral HBr molecules.
With no ions present in the benzene solution, it is electrically nonconductive.
PROBLEM 6.2.1.5
Consider the solutions presented:
(a) Which of the following sketches best represents the ions in a solution of Fe(NO3)3(aq)?
(b) Write a balanced chemical equation showing the products of the dissolution of Fe(NO3)3.
Answer a
Fe(NO3)3 is a strong electrolyte, thus it should completely dissociate into Fe3+ and −
(NO3 ) ions. Therefore, (z) best
represents the solution.
Answer b
3+ −
Fe(NO ) (s) ⟶ Fe (aq) + 3 NO (aq)
3 3 3
PROBLEM 6.2.1.6
What is the expected electrical conductivity of the following solutions?
a. NaOH(aq)
b. HCl(aq)
c. C6H12O6(aq) (glucose)
d. NH3(l)
Answer a
high conductivity (solute is an ionic compound that will dissociate when dissolved)
Answer b
high conductivity (solute is a strong acid and will ionize completely when dissolved)
Answer c
nonconductive (solute is a covalent compound, neither acid nor base, unreactive towards water)
Answer d
low conductivity (solute is a weak base and will partially ionize when dissolved)
PROBLEM 6.2.1.7
Why are most solid ionic compounds electrically nonconductive, whereas aqueous solutions of ionic compounds are good
conductors? Would you expect a liquid (molten) ionic compound to be electrically conductive or nonconductive? Explain.
6.2.1.2 https://chem.libretexts.org/@go/page/217283
Answer
A medium must contain freely mobile, charged entities to be electrically conductive. The ions present in a typical ionic solid
are immobilized in a crystalline lattice and so the solid is not able to support an electrical current. When the ions are
mobilized, either by melting the solid or dissolving it in water to dissociate the ions, current may flow and these forms of the
ionic compound are conductive.
PROBLEM 6.2.1.8
Suppose you are presented with a clear solution of sodium thiosulfate, Na2S2O3. How could you determine whether the solution
is unsaturated, saturated, or supersaturated?
Answer
Add a small crystal of N a S O . It will dissolve in an unsaturated solution, remain apparently unchanged in a saturated
2 2 3
PROBLEM 6.2.1.9
Refer to following figure for the following three questions:
a. How did the concentration of dissolved CO2 in the beverage change when the bottle was opened?
b. What caused this change?
c. Is the beverage unsaturated, saturated, or supersaturated with CO2?
Answer a
It decreased as some of the CO2 gas left the solution (evidenced by effervescence).
Answer b
Opening the bottle released the high-pressure CO2 gas above the beverage. The reduced CO2 gas pressure, per Henry’s law,
lowers the solubility for CO2.
Answer c
The dissolved CO2 concentration will continue to slowly decrease until equilibrium is reestablished between the beverage
and the very low CO2 gas pressure in the opened bottle. Immediately after opening, the beverage, therefore, contains
dissolved CO2 at a concentration greater than its solubility, a nonequilibrium condition, and is said to be supersaturated.
6.2.1.3 https://chem.libretexts.org/@go/page/217283
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
6.2.1: Practice Problems- Solutions Chemistry is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
6.2.1.4 https://chem.libretexts.org/@go/page/217283
6.3: Precipitation Reactions
Skills to Develop
Define precipitation reactions
Classify chemical reactions as one of these three types given appropriate descriptions or chemical equations
Write and balance precipitation equations in molecular, total ionic, and net ionic formats.
Faced with a wide range of varied interactions between chemical substances, scientists have found it convenient (or even
necessary) to classify chemical interactions by identifying common patterns of reactivity. The following three modules will provide
an introduction to three of the most prevalent types of chemical reactions: precipitation, acid-base, and oxidation-reduction.
group 1 metal cations (Li+, Na+, K+, Rb+, and Cs+) and ammonium
ion (NH )
+
the halide ions (Cl−, Br−, and I−) halides of Ag+, Hg , and Pb2+
2+
2
A vivid example of precipitation is observed when solutions of potassium iodide and lead nitrate are mixed, resulting in the
formation of solid lead iodide:
This observation is consistent with the solubility guidelines: The only insoluble compound among all those involved is lead iodide,
one of the exceptions to the general solubility of iodide salts.
The net ionic equation representing this reaction is:
2+ −
Pb (aq) + 2 I (aq) → PbI (s) (6.3.2)
2
6.3.1 https://chem.libretexts.org/@go/page/217279
Lead iodide is a bright yellow solid that was formerly used as an artist’s pigment known as iodine yellow (Figure 6.3.1 ). The
properties of pure PbI2 crystals make them useful for fabrication of X-ray and gamma ray detectors.
Video 6.3.1 : A precipitate of PbI2 forms when solutions containing Pb2+ and I− are mixed. Credit: MrLundScience
The solubility guidelines in Table 6.3.1 may be used to predict whether a precipitation reaction will occur when solutions of
soluble ionic compounds are mixed together. One merely needs to identify all the ions present in the solution and then consider if
possible cation/anion pairing could result in an insoluble compound. For example, mixing solutions of silver nitrate and sodium
fluoride will yield a solution containing Ag+, NO , Na+, and F− ions. Aside from the two ionic compounds originally present in
−
the solutions, AgNO3 and NaF, two additional ionic compounds may be derived from this collection of ions: NaNO3 and AgF. The
solubility guidelines indicate all nitrate salts are soluble but that AgF is one of the exceptions to the general solubility of fluoride
salts. A precipitation reaction, therefore, is predicted to occur, as described by the following equations:
NaF(aq) + AgNO (aq) → AgF(s) + NaNO (aq) (molecular) (6.3.3)
3 3
+ −
Ag (aq) + F (aq) → AgF(s) (net ionic) (6.3.4)
(b) The two possible products for this combination are LiC2H3O2 and AgCl. The solubility guidelines indicate AgCl is
insoluble, and so a precipitation reaction is expected. The net ionic equation for this reaction, derived in the manner detailed in
the previous module, is
+ −
Ag (aq) + Cl (aq) → AgCl(s)
(c) The two possible products for this combination are PbCO3 and NH4NO3. The solubility guidelines indicate PbCO3 is
insoluble, and so a precipitation reaction is expected. The net ionic equation for this reaction, derived in the manner detailed in
the previous module, is
2+ 2−
Pb (aq) + CO3 (aq) → PbCO (s)
3
6.3.2 https://chem.libretexts.org/@go/page/217279
Exercise 6.3.1
Which solution could be used to precipitate the barium ion, Ba2+, in a water sample: sodium chloride, sodium hydroxide, or
sodium sulfate? What is the formula for the expected precipitate?
Answer
sodium sulfate, BaSO4
There are two equivalent ways of considering a double-replacement equation: either the cations are swapped, or the anions are
swapped. (You cannot swap both; you would end up with the same substances you started with.) Either perspective should allow
you to predict the proper products, as long as you pair a cation with an anion and not a cation with a cation or an anion with an
anion.
Summary
6.3.3 https://chem.libretexts.org/@go/page/217279
Chemical reactions are classified according to similar patterns of behavior. A large number of important reactions are included in
three categories: precipitation, acid-base, and oxidation-reduction (redox). Precipitation reactions involve the formation of one or
more insoluble products.
Glossary
insoluble
of relatively low solubility; dissolving only to a slight extent
precipitate
insoluble product that forms from reaction of soluble reactants
precipitation reaction
reaction that produces one or more insoluble products; when reactants are ionic compounds, sometimes called double-
displacement or metathesis
double-displacement reaction
a type of reaction that occurs when the cations and anions switch between two reactants to form new products
soluble
of relatively high solubility; dissolving to a relatively large extent
solubility
the extent to which a substance may be dissolved in water, or any solvent
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
Types of Chemical Reactions Rachael Harper Delupio, Lost Angles Trade Technical College
Classifying Chemical Reactions,Nicola Burrmann, Heartland Community College
6.3: Precipitation Reactions is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
6.3.4 https://chem.libretexts.org/@go/page/217279
6.3.1: Practice Problems- Precipitation
PROBLEM 6.3.1.1
Aqueous hydrogen fluoride (hydrofluoric acid) is used to etch glass and to analyze minerals for their silicon content. Hydrogen
fluoride will also react with sand (silicon dioxide).
a. Write an equation for the reaction of solid silicon dioxide with hydrofluoric acid to yield gaseous silicon tetrafluoride and
liquid water.
b. The mineral fluorite (calcium fluoride) occurs extensively in Illinois. Solid calcium fluoride can also be prepared by the
reaction of aqueous solutions of calcium chloride and sodium fluoride, yielding aqueous sodium chloride as the other
product. Write complete and net ionic equations for this reaction.
Answer a
4 HF(aq) + SiO (s) → SiF (g) + 2 H O(l)
2 4 2
Answer b
complete ionic equation: 2 Na +
(aq) + 2 F
−
(aq) + Ca
2+
(aq) + 2 Cl
−
(aq) → CaF (s) + 2 Na
2
+
(aq) + 2 Cl
−
(aq)
PROBLEM 6.3.1.2
From the balanced molecular equations, write the complete ionic and net ionic equations for the following:
a. K C O (aq) + Ba(OH) (aq) → 2 KOH(aq) + BaC O (s)
2 2 4 2 2 2
Answer a
+ 2− 2+ − + −
2K (aq) + C O4 (aq) + Ba (aq) + 2 OH (aq) → 2 K (aq) + 2 OH (aq) + BaC O (s) (complete)
2 2 4
2+ 2−
Ba (aq) + C O (aq) → BaC O (s) (net)
2 4 2 4
Answer b
2+ − + 2− + −
Pb (aq) + 2 NO (aq) + 2 H (aq) + SO (aq) → PbSO (s) + 2 H (aq) + 2 NO (aq) (complete)
3 4 4 3
2+ 2−
Pb (aq) + SO 4 (aq) → PbSO (s) (net)
4
Answer c
+ 2−
CaCO (s) + 2 H (aq) + SO (aq) → CaSO (s) + CO (g) + H O(l) (complete)
3 4 4 2 2
+ 2−
CaCO (s) + 2 H (aq) + SO (aq) → CaSO (s) + CO (g) + H O(l) (net)
3 4 4 2 2
PROBLEM 6.3.1.3
Use the following equations to answer the next five questions:
i. H O(s) → H O(l)
2 2
3
(aq) → AgCl(s) + Na
+
(aq) + NO
−
3
(aq)
6.3.1.1 https://chem.libretexts.org/@go/page/217284
c. Which equation is not balanced?
d. Which is a net ionic equation?
Answer a
i. H 2 O(solid) → H2 O(liquid)
Answer b
iii.
Answer c
iii. 2 CH 3
OH(g) + 3 O (g) → 2 CO (g) + 4 H O(g)
2 2 2
Answer d
v.
PROBLEM 6.3.1.4
Write the molecular, total ionic, and net ionic equations for the following reactions:
a. Ca(OH) (aq) + HC H O (aq) →
2 2 3 2
Answer a
molecular: Ca(OH) 2
(aq) + 2 HC H O (aq) → Ca (C H O ) (aq) + 2 H O(l)
2 3 2 2 3 2 2 2
complete ionic: Ca 2+
(aq) + 2 OH
−
(aq) + 2 H
+
(aq) + 2 C H O
2 3
−
2
(aq) → Ca
2+
(aq) + 2 C H O
2 3
−
2
(aq) + 2 H O(l)
2
net ionic: 2 OH −
(aq) + 2 H
+
(aq) → 2 H O(l)
2
Answer b
molecular: 2 H 3
PO (aq) + 3 CaCl (aq) → Ca (PO ) (s) + 6 HCl(aq)
4 2 3 4 2
complete ionic: 6 H +
(aq) + 2 PO
3−
4
(aq) + 3 Ca
2+
(aq) + 6 Cl
−
(aq) → Ca (PO ) (s) + 6 H
3 4 2
+
(aq) + 6 Cl
−
(aq)
net ionic: 2 PO 3−
4
(aq) + 3 Ca
2+
(aq) → Ca (PO ) (s)
3 4 2
Problem 6.3.4
PROBLEM 6.3.1.5
Great Lakes Chemical Company produces bromine, Br2, from bromide salts such as NaBr, in Arkansas brine by treating the
brine with chlorine gas. Write a balanced equation for the reaction of NaBr with Cl2.
Answer
6.3.1.2 https://chem.libretexts.org/@go/page/217284
2 NaBr(aq) + Cl (g) → 2 NaCl(aq) + Br (l)
2 2
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
6.3.1: Practice Problems- Precipitation is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
6.3.1.3 https://chem.libretexts.org/@go/page/217284
6.4: Classifying Chemical Reactions (Acids and Bases)
Skills to Develop
Define three common types of chemical reactions (precipitation, acid-base, and oxidation-reduction)
Classify chemical reactions as one of these three types given appropriate descriptions or chemical equations
Identify common acids and bases
Acid-Base Reactions
The process represented by this equation confirms that hydrogen chloride is an acid. When dissolved in water, H3O+ ions are
produced by a chemical reaction in which H+ ions are transferred from HCl molecules to H2O molecules (Figure 6.4.1).
Figure 6.4.1 : When hydrogen chloride gas dissolves in water, (a) it reacts as an acid, transferring protons to water molecules to
yield (b) hydronium ions (and solvated chloride ions)
The nature of HCl is such that its reaction with water as just described is essentially 100% efficient: Virtually every HCl molecule
that dissolves in water will undergo this reaction. Acids that completely react in this fashion are called strong acids, and HCl is one
6.4.1 https://chem.libretexts.org/@go/page/271797
among just a handful of common acid compounds that are classified as strong (Table 6.4.1). A far greater number of compounds
behave as weak acids and only partially react with water, leaving a large majority of dissolved molecules in their original form and
generating a relatively small amount of hydronium ions. Weak acids are commonly encountered in nature, being the substances
partly responsible for the tangy taste of citrus fruits, the stinging sensation of insect bites, and the unpleasant smells associated with
body odor. A familiar example of a weak acid is acetic acid, the main ingredient in food vinegars:
− +
CH CO H(aq) + H O(l) ⇌ CH CO (aq) + H O (aq) (6.4.2)
3 2 2 3 2 3
When dissolved in water under typical conditions, only about 1% of acetic acid molecules are present in the ionized form,
CH CO
3
−
2
(Figure 6.4.2). (The use of a double-arrow in the equation above denotes the partial reaction aspect of this process, a
concept addressed fully in the chapters on chemical equilibrium.)
Figure 6.4.2 : (a) Fruits such as oranges, lemons, and grapefruit contain the weak acid citric acid. (b) Vinegars contain the weak
acid acetic acid. (credit a: modification of work by Scott Bauer; credit b: modification of work by Brücke-Osteuropa/Wikimedia
Commons)
Table 6.4.1: Common Strong Acids
Compound Formula Name in Aqueous Solution
HI hydroiodic acid
A base is a substance that will dissolve in water to yield hydroxide ions, OH−. The most common bases are ionic compounds
composed of alkali or alkaline earth metal cations (groups 1 and 2) combined with the hydroxide ion—for example, NaOH and
Ca(OH)2. When these compounds dissolve in water, hydroxide ions are released directly into the solution. For example, KOH and
Ba(OH)2 dissolve in water and dissociate completely to produce cations (K+ and Ba2+, respectively) and hydroxide ions, OH−.
These bases, along with other hydroxides that completely dissociate in water, are considered strong bases.
Consider as an example the dissolution of lye (sodium hydroxide) in water:
+ −
NaOH(s) → Na (aq) + OH (aq) (6.4.3)
This equation confirms that sodium hydroxide is a base. When dissolved in water, NaOH dissociates to yield Na and OH− ions. +
This is also true for any other ionic compound containing hydroxide ions. Since the dissociation process is essentially complete
when ionic compounds dissolve in water under typical conditions, NaOH and other ionic hydroxides are all classified as strong
bases.
6.4.2 https://chem.libretexts.org/@go/page/271797
Unlike ionic hydroxides, some compounds produce hydroxide ions when dissolved by chemically reacting with water molecules. In
all cases, these compounds react only partially and so are classified as weak bases. These types of compounds are also abundant in
nature and important commodities in various technologies. For example, global production of the weak base ammonia is typically
well over 100 metric tons annually, being widely used as an agricultural fertilizer, a raw material for chemical synthesis of other
compounds, and an active ingredient in household cleaners (Figure 6.4.3). When dissolved in water, ammonia reacts partially to
yield hydroxide ions, as shown here:
+ −
NH (aq) + H O(l) ⇌ NH (aq) + OH (aq) (6.4.4)
3 2 4
This is, by definition, an acid-base reaction, in this case involving the transfer of H+ ions from water molecules to ammonia
molecules. Under typical conditions, only about 1% of the dissolved ammonia is present as NH ions. +
Figure 6.4.3 : Ammonia is a weak base used in a variety of applications. (a) Pure ammonia is commonly applied as an
agricultural fertilizer. (b) Dilute solutions of ammonia are effective household cleansers. (credit a: modification of work by
National Resources Conservation Service; credit b: modification of work by pat00139)
The chemical reactions described in which acids and bases dissolved in water produce hydronium and hydroxide ions, respectively,
are, by definition, acid-base reactions. In these reactions, water serves as both a solvent and a reactant. A neutralization reaction is
a specific type of acid-base reaction in which the reactants are an acid and a base, the products are often a salt and water, and
neither reactant is the water itself:
acid + base → salt + water (6.4.5)
To illustrate a neutralization reaction, consider what happens when a typical antacid such as milk of magnesia (an aqueous
suspension of solid Mg(OH)2) is ingested to ease symptoms associated with excess stomach acid (HCl):
Mg (OH) (s) + 2 HCl(aq) → MgCl (aq) + 2 H O(l). (6.4.6)
2 2 2
Note that in addition to water, this reaction produces a salt, magnesium chloride.
A double-arrow is appropriate in this equation because it indicates the HOCl is a weak acid that has not reacted completely.
(b) The two reactants are provided, Ba(OH)2 and HNO3. Since this is a neutralization reaction, the two products will be water
and a salt composed of the cation of the ionic hydroxide (Ba2+) and the anion generated when the acid transfers its hydrogen ion
(NO ).
−
6.4.3 https://chem.libretexts.org/@go/page/271797
Exercise 6.4.1
Write the net ionic equation representing the neutralization of any strong acid with an ionic hydroxide. (Hint: Consider the ions
produced when a strong acid is dissolved in water.)
Answer
+ −
H O (aq) + OH (aq) → 2 H O(l)
3 2
Phet Simulation
Acid-Base Solutio
My Solutio
Introduction
6.4.4 https://chem.libretexts.org/@go/page/271797
excess acid in the esophagus by taking an antacid. As you may have guessed, antacids are bases. One of the most common
antacids is calcium carbonate, CaCO3. The reaction,
\ce{CaCO3}(𝑠)+\ce{2HCl}(𝑎𝑞)⇌\ce{CaCl2}(𝑎𝑞)+\ce{H2O}(𝑙)+\ce{CO2}(𝑔)
not only neutralizes stomach acid, it also produces CO2(g), which may result in a satisfying belch.
Milk of Magnesia is a suspension of the sparingly soluble base magnesium hydroxide, Mg(OH)2. It works according to the
reaction:
\ce{Mg(OH)2}(𝑠)⇌Mg2+(𝑎𝑞)+2OH−(𝑎𝑞)
The hydroxide ions generated in this equilibrium then go on to react with the hydronium ions from the stomach acid, so that:
H3O++OH−⇌2H2O(𝑙)
This reaction does not produce carbon dioxide, but magnesium-containing antacids can have a laxative effect. Several antacids
have aluminum hydroxide, Al(OH)3, as an active ingredient. The aluminum hydroxide tends to cause constipation, and some
antacids use aluminum hydroxide in concert with magnesium hydroxide to balance the side effects of the two substances.
6.4.5 https://chem.libretexts.org/@go/page/271797
Figure 6.4.4 : A neutralization reaction takes place between citric acid in lemons or acetic acid in vinegar, and the bases in the
flesh of fish.
Pickling is a method used to preserve vegetables using a naturally produced acidic environment. The vegetable, such as a
cucumber, is placed in a sealed jar submerged in a brine solution. The brine solution favors the growth of beneficial bacteria and
suppresses the growth of harmful bacteria. The beneficial bacteria feed on starches in the cucumber and produce lactic acid as a
waste product in a process called fermentation. The lactic acid eventually increases the acidity of the brine to a level that kills
any harmful bacteria, which require a basic environment. Without the harmful bacteria consuming the cucumbers they are able
to last much longer than if they were unprotected. A byproduct of the pickling process changes the flavor of the vegetables with
the acid making them taste sour.
Summary
Glossary
acid
substance that produces H3O+ when dissolved in water
acid-base reaction
reaction involving the transfer of a hydrogen ion between reactant species
base
substance that produces OH− when dissolved in water
salt
ionic compound that can be formed by the reaction of an acid with a base that contains a cation and an anion other than
hydroxide or oxide
strong acid
acid that reacts completely when dissolved in water to yield hydronium ions
strong base
base that reacts completely when dissolved in water to yield hydroxide ions
weak acid
6.4.6 https://chem.libretexts.org/@go/page/271797
acid that reacts only to a slight extent when dissolved in water to yield hydronium ions
weak base
base that reacts only to a slight extent when dissolved in water to yield hydroxide ions
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
TED-Ed’s commitment to creating lessons worth sharing is an extension of TED’s mission of spreading great ideas. Within
TED-Ed’s growing library of TED-Ed animations, you will find carefully curated educational videos, many of which represent
collaborations between talented educators and animators nominated through the TED-Ed website.
Feedback
6.4: Classifying Chemical Reactions (Acids and Bases) is shared under a not declared license and was authored, remixed, and/or curated by
LibreTexts.
6.4.7 https://chem.libretexts.org/@go/page/271797
6.4.1: Classifying Chemical Reactions (Acids and Bases) (Problems)
PROBLEM 6.4.1.1
Complete and balance the following acid-base equations:
a. HCl gas reacts with solid Ca(OH)2(s).
b. A solution of Sr(OH)2 is added to a solution of HNO3.
Answer a
2 HCl(g) + Ca (OH) (s) → CaCl (s) + 2 H O(l)
2 2 2
Answer b
Sr(OH) (aq) + 2 HNO (aq) → Sr(NO ) (aq) + 2 H O(l)
2 3 3 2 2
PROBLEM 6.4.1.2
Complete and balance the following acid-base equations:
a. A solution of HClO4 is added to a solution of LiOH.
b. Aqueous H2SO4 reacts with NaOH.
c. Ba(OH)2 reacts with HF gas.
Answer a
HClO (aq) + LiOH(aq) → LiClO (aq) + H O(l)
4 4 2
Answer b
H SO (aq) + 2 NaOH(aq) → Na SO (aq) + 2 H O(l)
2 4 2 4 2
Answer c
2 HF(aq) + Ba (OH) (aq) → BaF (aq) + 2 H O(l)
2 2 2
Problem 6.4.10
PROBLEM 6.4.1.3
Complete and balance the equations for the following acid-base neutralization reactions. If water is used as a solvent, write the
reactants and products as aqueous ions. In some cases, there may be more than one correct answer, depending on the amounts of
reactants used.
a. Mg (OH) (s) + HClO (aq) →
2 4
6.4.1.1 https://chem.libretexts.org/@go/page/271798
b. SO (g) + H
3 2
O(l) → (assume an excess of water and that the product dissolves)
c. SrO(s) + H 2
SO (l) →
4
Answer a
2+ −
Mg (OH) (s) + 2 HClO (aq) → Mg (aq) + 2 ClO (aq) + 2 H O(l)
2 4 4 2
Answer b
SO (g) + 2 H O(l) → H O
3 2 3
+
(aq) + HSO
−
4
(aq) , (a solution of H2SO4)
Answer c
SrO(s) + H SO (l) → SrSO (s) + H O
2 4 4 2
PROBLEM 6.4.1.4
Complete and balance the equations of the following reactions, each of which could be used to remove hydrogen sulfide from
natural gas:
a. Ca(OH) (s) + H 2 2
S(g) →
b. Na CO (aq) + H
2 3 2
S(g) →
Answer
Ca (OH) (s) + H S(g) → CaS(s) + 2 H O(l)
2 2 2
Answer
Na CO (aq) + H S(g) → Na S(aq) + CO (g) + H O(l)
2 3 2 2 2 2
PROBLEM 6.4.1.5
Calcium cyclamate Ca(C6H11NHSO3)2 is an artificial sweetener used in many countries around the world but is banned in the
United States. It can be purified industrially by converting it to the barium salt through reaction of the acid C6H11NHSO3H with
barium carbonate, treatment with sulfuric acid (barium sulfate is very insoluble), and then neutralization with calcium
hydroxide. Write the balanced equations for these reactions.
Answer
2C H NHSO H + BaCO → Ba (C H NHSO ) + H CO
6 11 3 3 6 11 3 2 2 3
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
6.4.1: Classifying Chemical Reactions (Acids and Bases) (Problems) is shared under a not declared license and was authored, remixed, and/or
curated by LibreTexts.
6.4.1.2 https://chem.libretexts.org/@go/page/271798
6.5: Classifying Chemical Reactions (Redox)
Skills to Develop
Define three common types of chemical reactions (precipitation, acid-base, and oxidation-reduction)
Classify chemical reactions as one of these three types given appropriate descriptions or chemical equations
Compute the oxidation states for elements in compounds
Oxidation-Reduction Reactions
Earth’s atmosphere contains about 20% molecular oxygen, O2, a chemically reactive gas that plays an essential role in the
metabolism of aerobic organisms and in many environmental processes that shape the world. The term oxidation was originally
used to describe chemical reactions involving O2, but its meaning has evolved to refer to a broad and important reaction class
known as oxidation-reduction (redox) reactions. A few examples of such reactions will be used to develop a clear picture of this
classification.
Some redox reactions involve the transfer of electrons between reactant species to yield ionic products, such as the reaction
between sodium and chlorine to yield sodium chloride:
2 Na(s) + Cl (g) → 2 NaCl(s) (6.5.1)
2
It is helpful to view the process with regard to each individual reactant, that is, to represent the fate of each reactant in the form of
an equation called a half-reaction:
+ −
2 Na(s) → 2 Na (s) + 2 e (6.5.2)
− −
Cl (g) + 2 e → 2 Cl (s) (6.5.3)
2
These equations show that Na atoms lose electrons while Cl atoms (in the Cl2 molecule) gain electrons, the “s” subscripts for the
resulting ions signifying they are present in the form of a solid ionic compound. For redox reactions of this sort, the loss and gain
of electrons define the complementary processes that occur:
oxidation = loss of electrons (6.5.4)
6.5.1 https://chem.libretexts.org/@go/page/271799
Figure 6.5.1 : Lion image by Tambako, reused under the Creative Commons Attribution 2.0 Generic license.
2. "Oil Rig"
Oxidation is loss
Reduction is gain
6.5.2 https://chem.libretexts.org/@go/page/271799
Figure 6.5.2 : Image by Peter Grima, reused under the Creative Commons Attribution 2.0 Generic license.
In this reaction, then, sodium is oxidized and chlorine undergoes reduction. Viewed from a more active perspective, sodium
functions as a reducing agent (reductant), since it provides electrons to (or reduces) chlorine. Likewise, chlorine functions as an
oxidizing agent (oxidant), as it effectively removes electrons from (oxidizes) sodium.
reducing agent = species that is oxidized (6.5.6)
Some redox processes, however, do not involve the transfer of electrons. Consider, for example, a reaction similar to the one
yielding NaCl:
H (g) + Cl (g) → 2 HCl(g) (6.5.8)
2 2
The product of this reaction is a covalent compound, so transfer of electrons in the explicit sense is not involved. To clarify the
similarity of this reaction to the previous one and permit an unambiguous definition of redox reactions, a property called oxidation
number has been defined. The oxidation number (or oxidation state) of an element in a compound is the charge its atoms would
possess if the compound was ionic. The following guidelines are used to assign oxidation numbers to each element in a molecule or
ion. These rules are hierarchical; if two rules conflict, the rule that is higher up on the list takes precedence:
1. The oxidation number of an atom in an elemental substance (a free element) is zero.
2. The oxidation number of a monatomic ion is equal to the ion’s charge.
3. The sum of oxidation numbers for all atoms in a molecule or polyatomic ion equals the charge on the molecule or ion.
4. Metals always have positive oxidation states when in a compound.
Group 1 (hydrogen) is ALWAYS +1
Group 2 (beryllium) is ALWAYS +2
5. Oxidation numbers for common nonmetals are usually assigned as follows:
Fluorine: always -1
Hydrogen: +1 when combined with nonmetals, −1 when combined with metals
1
Oxygen: −2 in most compounds, sometimes −1 (so-called peroxides, O 2−
2
), very rarely − (so-called superoxides, O ),
−
2
2
positive values when combined with F (values vary)
6.5.3 https://chem.libretexts.org/@go/page/271799
Halogens: −1 for other halogens except when combined with oxygen or other halogens (positive oxidation numbers in these
cases, varying values)
Group 6 (oxygen) is -2
Group 5 (nitrogen) is -3
Note: The proper convention for reporting charge is to write the number first, followed by the sign (e.g., 2+), while oxidation
number is written with the reversed sequence, sign followed by number (e.g., +2). This convention aims to emphasize the
distinction between these two related properties.
It is also important to note that it is possible to have fractional oxidation states.
c. Na2SO4
Solution
(a) According to guideline 1, the oxidation number for H is +1.
Using this oxidation number and the compound’s formula, guideline 4 may then be used to calculate the oxidation number for
sulfur:
charge on H S = 0 = (2 × +1) + (1 × x)
2
x = 0 − (2 × +1) = −2
x = −2 − (3 × −2) = +4
(c) For ionic compounds, it’s convenient to assign oxidation numbers for the cation and anion separately.
According to guideline 2, the oxidation number for sodium is +1.
Assuming the usual oxidation number for oxygen (−2 per guideline 3), the oxidation number for sulfur is calculated as directed
by guideline 4:
2−
charge on SO = −2 = (4 × −2) + (1 × x)
4
x = −2 − (4 × −2) = +6
Exercise 6.5.2
Assign oxidation states to the elements whose atoms are underlined in each of the following compounds or ions:
a. KNO3
b. AlH3
c. –
NH
–
+
4
d. H –
2PO
– 4
−
Answer a
N, +5
Answer b
Al, +3
Answer c
N, −3
6.5.4 https://chem.libretexts.org/@go/page/271799
Answer d
P, +5
Using the oxidation number concept, an all-inclusive definition of redox reaction has been established. Oxidation-reduction (redox)
reactions are those in which one or more elements involved undergo a change in oxidation number. While the vast majority of
redox reactions involve changes in oxidation number for two or more elements, a few interesting exceptions to this rule do exist as
shown below\). Definitions for the complementary processes of this reaction class are correspondingly revised as shown here:
oxidation = increase in oxidation number (6.5.9)
Returning to the reactions used to introduce this topic, they may now both be identified as redox processes. In the reaction between
sodium and chlorine to yield sodium chloride, sodium is oxidized (its oxidation number increases from 0 in Na to +1 in NaCl) and
chlorine is reduced (its oxidation number decreases from 0 in Cl2 to −1 in NaCl). In the reaction between molecular hydrogen and
chlorine, hydrogen is oxidized (its oxidation number increases from 0 in H2 to +1 in HCl) and chlorine is reduced (its oxidation
number decreases from 0 in Cl2 to −1 in HCl).
Several subclasses of redox reactions are recognized, including combustion reactions in which the reductant (also called a fuel) and
oxidant (often, but not necessarily, molecular oxygen) react vigorously and produce significant amounts of heat, and often light, in
the form of a flame. Solid rocket-fuel reactions such as the one depicted below are combustion processes. A typical propellant
reaction in which solid aluminum is oxidized by ammonium perchlorate is represented by this equation:
10 Al(s) + 6 NH ClO (s) → 4 Al O (s) + 2 AlCl (s) + 12 H O(g) + 3 N (g) (6.5.11)
4 4 2 3 3 2 2
Watch a brief video showing the test firing of a small-scale, prototype, hybrid rocket engine planned for use in the new Space
Launch System being developed by NASA. The first engines firing at 3 s (green flame) use a liquid fuel/oxidant mixture, and
the second, more powerful engines firing at 4 s (yellow flame) use a solid mixture.
Single-displacement (replacement) reactions are redox reactions in which an ion in solution is displaced (or replaced) via the
oxidation of a metallic element. One common example of this type of reaction is the acid oxidation of certain metals:
Metallic elements may also be oxidized by solutions of other metal salts; for example:
This reaction may be observed by placing copper wire in a solution containing a dissolved silver salt. Silver ions in solution are
reduced to elemental silver at the surface of the copper wire, and the resulting Cu2+ ions dissolve in the solution to yield a
characteristic blue color (Figure 6.5.6).
6.5.5 https://chem.libretexts.org/@go/page/271799
Figure 6.5.6 : (a) A copper wire is shown next to a solution containing silver(I) ions. (b) Displacement of dissolved silver ions by
copper ions results in (c) accumulation of gray-colored silver metal on the wire and development of a blue color in the solution,
due to dissolved copper ions. (credit: modification of work by Mark Ott)
Solution
Redox reactions are identified per definition if one or more elements undergo a change in oxidation number.
a. This is not a redox reaction, since oxidation numbers remain unchanged for all elements.
b. This is a redox reaction. Gallium is oxidized, its oxidation number increasing from 0 in Ga(l) to +3 in GaBr3(s). The
reducing agent is Ga(l). Bromine is reduced, its oxidation number decreasing from 0 in Br2(l) to −1 in GaBr3(s). The
oxidizing agent is Br2(l).
c. This is a redox reaction. It is a particularly interesting process, as it involves the same element, oxygen, undergoing both
oxidation and reduction (a so-called disproportionation reaction). Oxygen is oxidized, its oxidation number increasing from
−1 in H2O2(aq) to 0 in O2(g). Oxygen is also reduced, its oxidation number decreasing from −1 in H2O2(aq) to −2 in H2O(l).
For disproportionation reactions, the same substance functions as an oxidant and a reductant.
d. This is not a redox reaction, since oxidation numbers remain unchanged for all elements.
e. This is a redox reaction (combustion). Carbon is oxidized, its oxidation number increasing from −2 in C2H4(g) to +4 in
CO2(g). The reducing agent (fuel) is C2H4(g). Oxygen is reduced, its oxidation number decreasing from 0 in O2(g) to −2 in
H2O(l). The oxidizing agent is O2(g).
Exercise 6.5.3
This equation describes the production of tin(II) chloride:
Sn(s) + 2 HCl(g) → SnCl (s) + H (g) (6.5.14)
2 2
Is this a redox reaction? If so, provide a more specific name for the reaction if appropriate, and identify the oxidant and
reductant.
Answer
Yes, a single-replacement reaction. Sn(s)is the reductant, HCl(g) is the oxidant.
Summary
6.5.6 https://chem.libretexts.org/@go/page/271799
Redox Reactions: Crash Course Chemis…
Chemis…
Chemical reactions are classified according to similar patterns of behavior. A large number of important reactions are included in
three categories: precipitation, acid-base, and oxidation-reduction (redox). Precipitation reactions involve the formation of one or
more insoluble products. Acid-base reactions involve the transfer of hydrogen ions between reactants. Redox reactions involve a
change in oxidation number for one or more reactant elements. Writing balanced equations for some redox reactions that occur in
aqueous solutions is simplified by using a systematic approach called the half-reaction method.
Glossary
combustion reaction
vigorous redox reaction producing significant amounts of energy in the form of heat and, sometimes, light
oxidation
process in which an element’s oxidation number is increased by loss of electrons
oxidation-reduction reaction
(also, redox reaction) reaction involving a change in oxidation number for one or more reactant elements
oxidation number
(also, oxidation state) the charge each atom of an element would have in a compound if the compound were ionic
oxidizing agent
(also, oxidant) substance that brings about the oxidation of another substance, and in the process becomes reduced
reduction
process in which an element’s oxidation number is decreased by gain of electrons
reducing agent
(also, reductant) substance that brings about the reduction of another substance, and in the process becomes oxidized
single-displacement reaction
(also, replacement) redox reaction involving the oxidation of an elemental substance by an ionic species
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
6.5.7 https://chem.libretexts.org/@go/page/271799
Adelaide Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
6.5: Classifying Chemical Reactions (Redox) is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
6.5.8 https://chem.libretexts.org/@go/page/271799
6.5.1: Classifying Chemical Reactions (Redox) (Problems)
Some of these problems represent a culmination of all the reaction types we've covered in Unit 6!
PROBLEM 6.5.1.1
Indicate what type, or types, of reaction each of the following represents:
a. Ca(s) + Br (l) → CaBr (s)
2 2
Answer a
oxidation-reduction (addition)
Answer b
acid-base (neutralization)
Answer c
oxidation-reduction (combustion)
PROBLEM 6.5.1.2
Indicate what type, or types, of reaction each of the following represents:
a. H O(g) + C(s) → CO(g) + H (g)
2 2
Answer a
oxidation-reduction (single displacement)
Answer b
oxidation-reduction (dissociation)
Answer c
acid-base (neutralization)
Answer d
precipitation (double replacement)
PROBLEM 6.5.1.3
Silver can be separated from gold because silver dissolves in nitric acid while gold does not. Is the dissolution of silver in nitric
acid an acid-base reaction or an oxidation-reduction reaction? Explain your answer.
Answer
It is an oxidation-reduction reaction because the oxidation state of the silver changes during the reaction.
PROBLEM 6.5.1.4
Determine the oxidation states of the elements in the following compounds:
a. NaI
b. GdCl3
6.5.1.1 https://chem.libretexts.org/@go/page/271800
c. LiNO3
d. H2Se
e. Mg2Si
f. RbO2 (rubidium superoxide)
g. HF
Answer a
Na +1, I -1
Answer b
Gd +3, Cl -1
Answer c
Li +1, N +5, O -2
Answer d
H +1, Se -2
Answer e
Mg +2, Si -4
Answer f
Rb +1, O -1/2
Answer g
H +1, F -1
Problem 6.4.4
PROBLEM 6.5.1.5
Determine the oxidation states of the elements in the compounds listed. None of the oxygen-containing compounds are
peroxides or superoxides.
a. H3PO4
b. Al(OH)3
c. SeO2
d. KNO2
e. In2S3
f. P4O6
6.5.1.2 https://chem.libretexts.org/@go/page/271800
Answer a
H +1, P +5, O −2
Answer b
Al +3, H +1, O −2
Answer c
Se +4, O −2
Answer d
K +1, N +3, O −2
Answer e
In +3, S −2
Answer f
P +3, O −2
PROBLEM 6.5.1.6
Determine the oxidation states of the elements in the compounds listed. None of the oxygen-containing compounds are
peroxides or superoxides.
a. H2SO4
b. Ca(OH)2
c. BrOH
d. ClNO2
e. TiCl4
f. NaH
Answer a
H +1, O -2, S +6
Answer b
H +1, O -2, Ca +2
Answer c
H +1, O -2, Br +1,
Answer d
O -2, Cl -1, N +5
Answer e
Cl -1, Ti +4
Answer f
H +1, Na -1
PROBLEM 6.5.1.7
Classify the following as acid-base reactions or oxidation-reduction reactions:
a. Na S(aq) + 2 HCl(aq) → 2 NaCl(aq) + H S(g)
2 2
6.5.1.3 https://chem.libretexts.org/@go/page/271800
c. Mg(s) + Cl (g) → MgCl (s)2 2
Answer a
acid-base
Answer b
oxidation-reduction: Na is oxidized, H+ is reduced
Answer c
oxidation-reduction: Mg is oxidized, Cl2 is reduced
Answer d
acid-base
Answer e
oxidation-reduction: P3− is oxidized, O2 is reduced
Answer f
acid-base
PROBLEM 6.5.1.8
Identify the atoms that are oxidized and reduced, the change in oxidation state for each, and the oxidizing and reducing agents in
each of the following equations:
a. Mg(s) + NiCl (aq) → MgCl (aq) + Ni(s)
2 2
Answer a
Mg: 0 → +2; loses electrons; oxidized; reducing agent
Ni: +2 → 0; gains electrons; reduced; oxidizing agent
Answer b
P: +3 → +5; loses electrons; oxidized; reducing agent
Cl: 0 → -1; gains electrons; reduced; oxidizing agent
Answer c
C: -2 → +4; loses electrons; oxidized; reducing agent
O: 0 → -2; gains electrons; reduced; oxidizing agent
Answer d
Zn: 0 → +2; loses electrons; oxidized; reducing agent
H: +1 → 0; gains electrons; reduced; oxidizing agent
Answer e
6.5.1.4 https://chem.libretexts.org/@go/page/271800
S: +2 → +5/2; loses electrons; oxidized; reducing agent
I: 0 → -1; gains electrons; reduced; oxidizing agent
Answer f
Cu: 0 → +2; loses electrons; oxidized; reducing agent
N: +5 → +2; gains electrons; reduced; oxidizing agent
Problem 6.4.8
PROBLEM 6.5.1.10
When heated to 700–800 °C, diamonds, which are pure carbon, are oxidized by atmospheric oxygen. (They burn!) Write the
balanced equation for this reaction.
Answer
C (s) + O (g) → CO (g)
diamond 2 2
PROBLEM 6.5.1.11
The military has experimented with lasers that produce very intense light when fluorine combines explosively with hydrogen.
What is the balanced equation for this reaction?
Answer
H (g) + F (g) → 2 HF(g)
2 2
PROBLEM 6.5.1.12
In a common experiment in the general chemistry laboratory (that you will do if you haven't already), magnesium metal is
heated in air to produce MgO. MgO is a white solid, but in these experiments it often looks gray, due to small amounts of
Mg3N2, a compound formed as some of the magnesium reacts with nitrogen. Write a balanced equation for each reaction.
Answer
2 Mg(s) + O (g) → 2 MgO(s)
2
PROBLEM 6.5.1.13
Copper(II) sulfide is oxidized by molecular oxygen to produce gaseous sulfur trioxide and solid copper(II) oxide. The gaseous
product then reacts with liquid water to produce liquid hydrogen sulfate as the only product. Write the two balanced equations
which represent these reactions.
6.5.1.5 https://chem.libretexts.org/@go/page/271800
Answer
CuS(s) + 2 O (g) → SO (g) + CuO(s)
2 3
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
6.5.1: Classifying Chemical Reactions (Redox) (Problems) is shared under a not declared license and was authored, remixed, and/or curated by
LibreTexts.
6.5.1.6 https://chem.libretexts.org/@go/page/271800
CHAPTER OVERVIEW
7: Ideal Gas Behavior
Unit 7 Objectives
By the end of this unit, you will be able to:
Define the property of temperature and pressure
Define and convert among the units of temperature and pressure
Describe the operation of common tools for measuring gas pressure
Identify the mathematical relationships between the various properties of gases
Use the ideal gas law, and related gas laws, to compute the values of various gas properties under specified conditions
Use the ideal gas law to compute gas densities and molar masses
Perform stoichiometric calculations involving gaseous substances
State Dalton’s law of partial pressures and use it in calculations involving gaseous mixtures
State the postulates of the kinetic-molecular theory
Use this theory’s postulates to explain the gas laws
7: Ideal Gas Behavior is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
1
7.1: Temperature and Pressure
Skills to Develop
Define the property of temperature and pressure
Define and convert among the units of temperature and pressure
Describe the operation of common tools for measuring gas pressure
Temperature
Video 7.1.1 : An introduction to temperature in terms of the Ideal Gas Law, which we will cover later in this Unit.
Temperature is an intensive property. The SI unit of temperature is the kelvin (K). The IUPAC convention is to use kelvin (all
lowercase) for the word, K (uppercase) for the unit symbol, and neither the word “degree” nor the degree symbol (°). The degree
Celsius (°C) is also allowed in the SI system, with both the word “degree” and the degree symbol used for Celsius measurements.
Celsius degrees are the same magnitude as those of kelvin, but the two scales place their zeros in different places. Water freezes at
273.15 K (0 °C) and boils at 373.15 K (100 °C) by definition, and normal human body temperature is approximately 310 K (37
°C).
where
y = length in feet,
7.1.1 https://chem.libretexts.org/@go/page/217288
x = length in inches, and
the proportionality constant, m, is the conversion factor.
The Celsius and Fahrenheit temperature scales, however, do not share a common zero point, and so the relationship between these
two scales is a linear one rather than a proportional one (y = mx + b ). Consequently, converting a temperature from one of these
scales into the other requires more than simple multiplication by a conversion factor, m, it also must take into account differences
in the scales’ zero points (b ).
The linear equation relating Celsius and Fahrenheit temperatures is easily derived from the two temperatures used to define each
scale. Representing the Celsius temperature as x and the Fahrenheit temperature as y , the slope, m, is computed to be:
Δy
m = (7.1.2)
Δx
∘ ∘
212 F − 32 F
= (7.1.3)
∘ ∘
100 C −0 C
∘
180 F
= (7.1.4)
∘
100 C
∘
9 F
= (7.1.5)
∘
5 C
The y-intercept of the equation, b, is then calculated using either of the equivalent temperature pairs, (100 °C, 212 °F) or (0 °C, 32
°F), as:
b = y − mx (7.1.6)
∘
9 F
∘ ∘
= 32 F− ×0 C (7.1.7)
∘
5 C
∘
= 32 F (7.1.8)
An abbreviated form of this equation that omits the measurement units is:
9
T∘ F = × T∘ C + 32 (7.1.10)
5
Rearrangement of this equation yields the form useful for converting from Fahrenheit to Celsius:
5
T∘ C = (T∘ F + 32) (7.1.11)
9
As mentioned earlier in this chapter, the SI unit of temperature is the kelvin (K). Unlike the Celsius and Fahrenheit scales, the
kelvin scale is an absolute temperature scale in which 0 (zero) K corresponds to the lowest temperature that can theoretically be
achieved. The early 19th-century discovery of the relationship between a gas's volume and temperature suggested that the volume
of a gas would be zero at −273.15 °C. In 1848, British physicist William Thompson, who later adopted the title of Lord Kelvin,
proposed an absolute temperature scale based on this concept (further treatment of this topic is provided in this text’s chapter on
gases).
The freezing temperature of water on this scale is 273.15 K and its boiling temperature 373.15 K. Notice the numerical difference
in these two reference temperatures is 100, the same as for the Celsius scale, and so the linear relation between these two
K
temperature scales will exhibit a slope of 1 ∘
. Following the same approach, the equations for converting between the kelvin and
C
Celsius temperature scales are derived to be:
TK = T∘ C + 273.15 (7.1.12)
T∘ C = TK − 273.15 (7.1.13)
7.1.2 https://chem.libretexts.org/@go/page/217288
The 273.15 in these equations has been determined experimentally, so it is not exact. Figure 7.1.1 shows the relationship among the
three temperature scales. Recall that we do not use the degree sign with temperatures on the kelvin scale.
Figure 7.1.1 : The Fahrenheit, Celsius, and kelvin temperature scales are compared.
Although the kelvin (absolute) temperature scale is the official SI temperature scale, Celsius is commonly used in many scientific
contexts and is the scale of choice for nonscience contexts in almost all areas of the world. Very few countries (the U.S. and its
territories, the Bahamas, Belize, Cayman Islands, and Palau) still use Fahrenheit for weather, medicine, and cooking.
9 9
∘ ∘ ∘
F = C + 32.0 = ( × 37.0) + 32.0 = 66.6 + 32.0 = 98.6 F
5 5
Exercise 7.1.1
Convert 80.92 °C to K and °F.
Answer
354.07 K, 177.7 °F
∘ 2
K= C + 273.15 = 230 + 273 = 503 K → 5.0 × 10 K (two signif icant f igures)
7.1.3 https://chem.libretexts.org/@go/page/217288
Exercise 7.1.2
Convert 50 °F to °C and K.
Answer
10 °C, 280 K
Pressure
The earth’s atmosphere exerts a pressure, as does any other gas. Although we do not normally notice atmospheric pressure, we are
sensitive to pressure changes—for example, when your ears “pop” during take-off and landing while flying, or when you dive
underwater. Gas pressure is caused by the force exerted by gas molecules colliding with the surfaces of objects (Figure 7.1.2).
Although the force of each collision is very small, any surface of appreciable area experiences a large number of collisions in a
short time, which can result in a high pressure. In fact, normal air pressure is strong enough to crush a metal container when not
balanced by equal pressure from inside the container.
Figure 7.1.2 : The atmosphere above us exerts a large pressure on objects at the surface of the earth, roughly equal to the weight
of a bowling ball pressing on an area the size of a human thumbnail.
Atmospheric pressure is caused by the weight of the column of air molecules in the atmosphere above an object, such as the tanker
car. At sea level, this pressure is roughly the same as that exerted by a full-grown African elephant standing on a doormat, or a
typical bowling ball resting on your thumbnail. These may seem like huge amounts, and they are, but life on earth has evolved
under such atmospheric pressure. If you actually perch a bowling ball on your thumbnail, the pressure experienced is twice the
usual pressure, and the sensation is unpleasant.
7.1.4 https://chem.libretexts.org/@go/page/217288
Railroad tank car vacuum implosion
Video 7.1.2 : A dramatic illustration of atmospheric pressure is provided in this brief video, which shows a railway tanker car
imploding when its internal pressure is decreased.
7.1.5 https://chem.libretexts.org/@go/page/217288
Crush a 55 gallon drum with air pressure
Since pressure is directly proportional to force and inversely proportional to area (Equation 7.1.14), pressure can be increased
either by either increasing the amount of force or by decreasing the area over which it is applied. Correspondingly, pressure
can be decreased by either decreasing the force or increasing the area.
Let’s apply the definition of pressure (Equation 7.1.14) to determine which would be more likely to fall through thin ice in Figure
7.1.3.—the elephant or the figure skater?
7.1.6 https://chem.libretexts.org/@go/page/217288
Figure 7.1.3 : Although (a) an elephant’s weight is large, creating a very large force on the ground, (b) the figure skater exerts a
much higher pressure on the ice due to the small surface area of her skates. (credit a: modification of work by Guido da Rozze;
credit b: modification of work by Ryosuke Yagi).
A large African elephant can weigh 7 tons, supported on four feet, each with a diameter of about 1.5 ft (footprint area of 250 in2),
so the pressure exerted by each foot is about 14 lb/in2:
lb 1 elephant 1 f oot 2
pressure per elephant f oot = 14, 000 × × = 14 lb/i n (7.1.15)
elephant 4 f eet 250 in2
The figure skater weighs about 120 lbs, supported on two skate blades, each with an area of about 2 in2, so the pressure exerted by
each blade is about 30 lb/in2:
lb 1 skater 1 blade
2
pressure per skate blade = 120 × × = 30 lb/i n (7.1.16)
2
skater 2 blades 2 in
Even though the elephant is more than one hundred-times heavier than the skater, it exerts less than one-half of the pressure and
would therefore be less likely to fall though thin ice. On the other hand, if the skater removes her skates and stands with bare feet
(or regular footwear) on the ice, the larger area over which her weight is applied greatly reduces the pressure exerted:
lb 1 skater 1 f oot 2
pressure per human f oot = 120 × × = 2 lb/i n (7.1.17)
2
skater 2 f eet 30 in
The SI unit of pressure is the pascal (Pa), with 1 Pa = 1 N/m2, where N is the newton, a unit of force defined as 1 kg m/s2. One
pascal is a small pressure; in many cases, it is more convenient to use units of kilopascal (1 kPa = 1000 Pa) or bar (1 bar = 100,000
Pa). In the United States, pressure is often measured in pounds of force on an area of one square inch—pounds per square inch (psi)
—for example, in car tires. Pressure can also be measured using the unit atmosphere (atm), which originally represented the
average sea level air pressure at the approximate latitude of Paris (45°). Table 7.1.1 provides some information on these and a few
other common units for pressure measurements
Table 7.1.1: Pressure Units
Unit Name and Abbreviation Definition or Relation to Other Unit Comment
pounds per square inch (psi) air pressure at sea level is ~14.7 psi
7.1.7 https://chem.libretexts.org/@go/page/217288
Example 7.1.3 : Conversion of Pressure Units
The United States National Weather Service reports pressure in both inches of Hg and millibars. Convert a pressure of 29.2 in.
Hg into:
a. torr
b. atm
c. kPa
d. mbar
Solution
This is a unit conversion problem. The relationships between the various pressure units are given in Table 7.1.1.
25.4 mm 1 torr
a. 29.2 in Hg × × = 742 torr
1 in 1 mm Hg
1 atm
b. 742 torr × = 0.976 atm
760 torr
101.325 kPa
c. 742 torr × = 98.9 kPa
760 torr
Exercise 7.1.3
A typical barometric pressure in Kansas City is 740 torr. What is this pressure in atmospheres, in millimeters of mercury, in
kilopascals, and in bar?
Answer
0.974 atm; 740 mm Hg; 98.7 kPa; 0.987 bar
We can measure atmospheric pressure, the force exerted by the atmosphere on the earth’s surface, with a barometer (Figure 7.1.4).
A barometer is a glass tube that is closed at one end, filled with a nonvolatile liquid such as mercury, and then inverted and
immersed in a container of that liquid. The atmosphere exerts pressure on the liquid outside the tube, the column of liquid exerts
pressure inside the tube, and the pressure at the liquid surface is the same inside and outside the tube. The height of the liquid in the
tube is therefore proportional to the pressure exerted by the atmosphere.
7.1.8 https://chem.libretexts.org/@go/page/217288
Figure 7.1.4 : In a barometer, the height, h, of the column of liquid is used as a measurement of the air pressure. Using very dense
liquid mercury (left) permits the construction of reasonably sized barometers, whereas using water (right) would require a
barometer more than 30 feet tall.
where
h is the height of the fluid,
ρ is the density of the fluid, and
A manometer is a device similar to a barometer that can be used to measure the pressure of a gas trapped in a container. A closed-
end manometer is a U-shaped tube with one closed arm, one arm that connects to the gas to be measured, and a nonvolatile liquid
(usually mercury) in between. As with a barometer, the distance between the liquid levels in the two arms of the tube (h in the
diagram) is proportional to the pressure of the gas in the container. An open-end manometer (Figure 7.1.5) is the same as a closed-
end manometer, but one of its arms is open to the atmosphere. In this case, the distance between the liquid levels corresponds to the
difference in pressure between the gas in the container and the atmosphere.
7.1.9 https://chem.libretexts.org/@go/page/217288
Figure 7.1.5 : A manometer can be used to measure the pressure of a gas. The (difference in) height between the liquid levels (h)
is a measure of the pressure. Mercury is usually used because of its large density.
Application: Measuring Blood Pressure
Blood pressure is measured using a device called a sphygmomanometer (Greek sphygmos = “pulse”). It consists of an inflatable
cuff to restrict blood flow, a manometer to measure the pressure, and a method of determining when blood flow begins and
when it becomes impeded (Figure 7.1.6). Since its invention in 1881, it has been an essential medical device. There are many
types of sphygmomanometers: manual ones that require a stethoscope and are used by medical professionals; mercury ones,
used when the most accuracy is required; less accurate mechanical ones; and digital ones that can be used with little training but
that have limitations. When using a sphygmomanometer, the cuff is placed around the upper arm and inflated until blood flow is
completely blocked, then slowly released. As the heart beats, blood forced through the arteries causes a rise in pressure. This
rise in pressure at which blood flow begins is the systolic pressure—the peak pressure in the cardiac cycle. When the cuff’s
pressure equals the arterial systolic pressure, blood flows past the cuff, creating audible sounds that can be heard using a
stethoscope. This is followed by a decrease in pressure as the heart’s ventricles prepare for another beat. As cuff pressure
continues to decrease, eventually sound is no longer heard; this is the diastolic pressure—the lowest pressure (resting phase) in
the cardiac cycle. Blood pressure units from a sphygmomanometer are in terms of millimeters of mercury (mm Hg).
Figure 7.1.6 : (a) A medical technician prepares to measure a patient’s blood pressure with a sphygmomanometer. (b) A typical
sphygmomanometer uses a valved rubber bulb to inflate the cuff and a diaphragm gauge to measure pressure. (credit a:
modification of work by Master Sgt. Jeffrey Allen)
7.1.10 https://chem.libretexts.org/@go/page/217288
Figure 7.1.7 : Meteorologists use weather maps to describe and predict weather. Regions of high (H) and low (L) pressure
have large effects on weather conditions. The gray lines represent locations of constant pressure known as isobars. (credit:
modification of work by National Oceanic and Atmospheric Administration)
In terms of weather, low-pressure systems occur when the earth’s surface atmospheric pressure is lower than the surrounding
environment: Moist air rises and condenses, producing clouds. Movement of moisture and air within various weather fronts
instigates most weather events.
The atmosphere is the gaseous layer that surrounds a planet. Earth’s atmosphere, which is roughly 100–125 km thick, consists of
roughly 78.1% nitrogen and 21.0% oxygen, and can be subdivided further into the regions shown in Figure 7.1.8: the exosphere
(furthest from earth, > 700 km above sea level), the thermosphere (80–700 km), the mesosphere (50–80 km), the stratosphere
(second lowest level of our atmosphere, 12–50 km above sea level), and the troposphere (up to 12 km above sea level, roughly
80% of the earth’s atmosphere by mass and the layer where most weather events originate). As you go higher in the troposphere,
air density and temperature both decrease.
Figure 7.1.8 : Earth’s atmosphere has five layers: the troposphere, the stratosphere, the mesosphere, the thermosphere, and the
exosphere.
Climatology is the study of the climate, averaged weather conditions over long time periods, using atmospheric data. However,
climatologists study patterns and effects that occur over decades, centuries, and millennia, rather than shorter time frames of
hours, days, and weeks like meteorologists. Atmospheric science is an even broader field, combining meteorology, climatology,
and other scientific disciplines that study the atmosphere.
Summary
7.1.11 https://chem.libretexts.org/@go/page/217288
Key Equations
5
T∘ C = × T∘ F − 32
9
9
T∘ F = × T∘ C + 32
5
∘
TK = C + 273.15
T∘ C = K − 273.15
F
P =
A
p = hρg
Glossary
atmosphere (atm)
unit of pressure; 1 atm = 101,325 Pa
bar
(bar or b) unit of pressure; 1 bar = 100,000 Pa
barometer
device used to measure atmospheric pressure
Celsius (°C)
unit of temperature; water freezes at 0 °C and boils at 100 °C on this scale
Fahrenheit
unit of temperature; water freezes at 32 °F and boils at 212 °F on this scale
hydrostatic pressure
pressure exerted by a fluid due to gravity
kelvin (K)
SI unit of temperature; 273.15 K = 0 ºC
manometer
device used to measure the pressure of a gas trapped in a container
pascal (Pa)
SI unit of pressure; 1 Pa = 1 N/m2
pressure
force exerted per unit area
torr
1
unit of pressure; 1 torr = atm
760
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Techonology
Crash Course Physics: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
7.1.12 https://chem.libretexts.org/@go/page/217288
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
TED-Ed’s commitment to creating lessons worth sharing is an extension of TED’s mission of spreading great ideas. Within
TED-Ed’s growing library of TED-Ed animations, you will find carefully curated educational videos, many of which represent
collaborations between talented educators and animators nominated through the TED-Ed website.
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
7.1: Temperature and Pressure is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
7.1.13 https://chem.libretexts.org/@go/page/217288
7.1.1: Practice Problems- Temperature and Pressure
PROBLEM 7.1.1.1
Convert the boiling temperature of gold, 2966 °C, into degrees Fahrenheit and kelvin.
Answer
5371 °F
3239 K
PROBLEM 7.1.1.2
Convert the temperature of scalding water, 54 °C, into degrees Fahrenheit and kelvin.
Answer
129.2 °F
327 K
Problem 7.1.2
PROBLEM 7.1.1.3
Convert the temperature of the coldest area in a freezer, −10 °F, to degrees Celsius and kelvin.
Answer
−23 °C
250 K
PROBLEM 7.1.1.4
Convert the temperature of dry ice, −77 °C, into degrees Fahrenheit and kelvin.
Answer
-106.6 °F
196 K
7.1.1.1 https://chem.libretexts.org/@go/page/217289
Problem 7.1.4
PROBLEM 7.1.1.5
Convert the boiling temperature of liquid ammonia, −28.1 °F, into degrees Celsius and kelvin.
Answer
−33.4 °C
239.8 K
PROBLEM 7.1.1.6
The label on a pressurized can of spray disinfectant warns against heating the can above 130 °F. What are the corresponding
temperatures on the Celsius and kelvin temperature scales?
Answer
90 °C
363 K
Problem 7.1.6
PROBLEM 7.1.1.7
The weather in Europe was unusually warm during the summer of 1995. The TV news reported temperatures as high as 45 °C.
What was the temperature on the Fahrenheit scale?
Answer
113 °F
7.1.1.2 https://chem.libretexts.org/@go/page/217289
PROBLEM 7.1.1.8
Why do some small bridges have weight limits that depend on how many wheels or axles the crossing vehicle has?
Answer
Pressure is defined as force per unit area. Similarly to the example in the text of the elephant and the figure skater, the more
wheels or axles on the vehicle, the more area the weight is spread over, causing less pressure to be exerted on any one point
on the bridge.
PROBLEM 7.1.1.9
A typical barometric pressure in Redding, California, is about 750 mm Hg. Calculate this pressure in atm and kPa
Answer
0.987 atm
99.97 kPa
PROBLEM 7.1.1.10
A typical barometric pressure in Denver, Colorado, is 615 mm Hg. What is this pressure in atmospheres and kilopascals?
Answer
0.809 atm
82.0 kPa
Problem 7.1.10
PROBLEM 7.1.1.11
Canadian tire pressure gauges are marked in units of kilopascals. What reading on such a gauge corresponds to 32 psi?
Answer
2.2 × 102 kPa
PROBLEM 7.1.1.12
During the Viking mission landings on Mars, the atmospheric pressure was determined to be on the average about 6.50 millibars
(1 bar = 0.987 atm). What is that pressure in torr and kPa?
7.1.1.3 https://chem.libretexts.org/@go/page/217289
Answer
4.88 torr
0.650 kPa
Problem 7.1.12
PROBLEM 7.1.1.13
The pressure of the atmosphere on the surface of the planet Venus is about 88.8 atm. Compare that pressure in psi to the normal
pressure on earth at sea level in psi.
Answer
Earth: 14.7 psi
Venus: 13.1 × 103 psi
PROBLEM 7.1.1.14
Consider this scenario and answer the following questions: On a mid-August day in the northeastern United States, the
following information appeared in the local newspaper: atmospheric pressure at sea level 29.97 in. Hg, 1013.9 mbar.
a. What was the pressure in kPa?
b. The pressure near the seacoast in the northeastern United States is usually reported near 30.0 in. Hg. During a hurricane, the
pressure may fall to near 28.0 in. Hg. Calculate the drop in pressure in torr.
Answer a
101.5 kPa
Answer b
51 torr drop
PROBLEM 7.1.1.15
The pressure of a sample of gas is measured at sea level is 264 mmHg. Determine the pressure of the gas in:
a. torr
b. Pa
7.1.1.4 https://chem.libretexts.org/@go/page/217289
c. bar
Answer a
264 torr
Answer b
35197 Pa
Answer c
0.352 bar
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
7.1.1: Practice Problems- Temperature and Pressure is shared under a not declared license and was authored, remixed, and/or curated by
LibreTexts.
7.1.1.5 https://chem.libretexts.org/@go/page/217289
7.2: The Gas Laws
Skills to Develop
Identify the mathematical relationships between the various properties of gases
Use the ideal gas law, and related gas laws, to compute the values of various gas properties under specified conditions
During the seventeenth and especially eighteenth centuries, driven both by a desire to understand nature and a quest to make
balloons in which they could fly (Figure 7.2.1), a number of scientists established the relationships between the macroscopic
physical properties of gases, that is, pressure, volume, temperature, and amount of gas. Although their measurements were not
precise by today’s standards, they were able to determine the mathematical relationships between pairs of these variables (e.g.,
pressure and temperature, pressure and volume) that hold for an ideal gas—a hypothetical construct that real gases approximate
under certain conditions. Eventually, these individual laws were combined into a single equation—the ideal gas law—that relates
gas quantities for gases and is quite accurate for low pressures and moderate temperatures. We will consider the key developments
in individual relationships (for pedagogical reasons not quite in historical order), then put them together in the ideal gas law.
Figure 7.2.1 : In 1783, the first (a) hydrogen-filled balloon flight, (b) manned hot air balloon flight, and (c) manned hydrogen-
filled balloon flight occurred. When the hydrogen-filled balloon depicted in (a) landed, the frightened villagers of Gonesse
reportedly destroyed it with pitchforks and knives. The launch of the latter was reportedly viewed by 400,000 people in Paris.
Figure 7.2.2 : The effect of temperature on gas pressure: When the hot plate is off, the pressure of the gas in the sphere is
relatively low. As the gas is heated, the pressure of the gas in the sphere increases.
This relationship between temperature and pressure is observed for any sample of gas confined to a constant volume. An example
of experimental pressure-temperature data is shown for a sample of air under these conditions in Figure 7.2.3. We find that
temperature and pressure are linearly related, and if the temperature is on the kelvin scale, then P and T are directly proportional
7.2.1 https://chem.libretexts.org/@go/page/217290
(again, when volume and moles of gas are held constant); if the temperature on the kelvin scale increases by a certain factor, the
gas pressure increases by the same factor.
Figure 7.2.3 : For a constant volume and amount of air, the pressure and temperature are directly proportional, provided the
temperature is in kelvin. (Measurements cannot be made at lower temperatures because of the condensation of the gas.) When this
line is extrapolated to lower pressures, it reaches a pressure of 0 at –273 °C, which is 0 on the kelvin scale and the lowest possible
temperature, called absolute zero.
Guillaume Amontons was the first to empirically establish the relationship between the pressure and the temperature of a gas
(~1700), and Joseph Louis Gay-Lussac determined the relationship more precisely (~1800). Because of this, the P-T relationship
for gases is known as either Amontons’s law or Gay-Lussac’s law. Under either name, it states that the pressure of a given amount
of gas is directly proportional to its temperature on the kelvin scale when the volume is held constant. Mathematically, this can be
written:
P ∝ T or P = constant × T or P = k × T (7.2.1)
where ∝ means “is proportional to,” and k is a proportionality constant that depends on the identity, amount, and volume of the gas.
P P
For a confined, constant volume of gas, the ratio is therefore constant (i.e., =k ). If the gas is initially in “Condition 1” (with
T T
P1 P2
P = P1 and T = T1), and then changes to “Condition 2” (with P = P2 and T = T2), we have that =k and =k , which
T1 T2
P1 P2
reduces to = . This equation is useful for pressure-temperature calculations for a confined gas at constant volume. Note
T1 T2
that temperatures must be on the kelvin scale for any gas law calculations (0 on the kelvin scale and the lowest possible
temperature is called absolute zero). (Also note that there are at least three ways we can describe how the pressure of a gas changes
as its temperature changes: We can use a table of values, a graph, or a mathematical equation.)
7.2.2 https://chem.libretexts.org/@go/page/217290
P1 P2 360 kPa P2
= which means that =
T1 T2 297 K 323 K
Exercise 7.2.1
A sample of nitrogen, N2, occupies 45.0 mL at 27 °C and 600 torr. What pressure will it have if cooled to –73 °C while the
volume remains constant?
Answer
400 torr
Video 7.2.1 : This video shows how cooling and heating a gas causes its volume to decrease or increase, respectively.
These examples of the effect of temperature on the volume of a given amount of a confined gas at constant pressure are true in
general: The volume increases as the temperature increases, and decreases as the temperature decreases. Volume-temperature data
for a 1-mole sample of methane gas at 1 atm are listed and graphed in Figure 7.2.2.
7.2.3 https://chem.libretexts.org/@go/page/217290
Figure 7.2.4 : The volume and temperature are linearly related for 1 mole of methane gas at a constant pressure of 1 atm. If the
temperature is in kelvin, volume and temperature are directly proportional. The line stops at 111 K because methane liquefies at
this temperature; when extrapolated, it intersects the graph’s origin, representing a temperature of absolute zero.
The relationship between the volume and temperature of a given amount of gas at constant pressure is known as Charles’s law in
recognition of the French scientist and balloon flight pioneer Jacques Alexandre César Charles. Charles’s law states that the volume
of a given amount of gas is directly proportional to its temperature on the kelvin scale when the pressure is held constant.
Mathematically, this can be written as:
V αT or V = constant ⋅ T or V = k ⋅ T or V1 / T1 = V2 / T2 (7.2.2)
with k being a proportionality constant that depends on the amount and pressure of the gas.
V
For a confined, constant pressure gas sample, is constant (i.e., the ratio = k), and as seen with the P-T relationship, this leads to
T
V1 V2
another form of Charles’s law: = .
T1 T2
Figure 7.2.5 : An illustration of Charles's Law, showing that as volume decreases, so does temperature, and vice versa. This file is
in the public domain because it was created by NASA . NASA's copyright policy states that "NASA material is not
copyrighted unless otherwise specified ". ( NASA copyright policy and JPL Image Use Policy ).
0.300 L × 303 K
Rearranging and solving gives: V 2 = = 0.321 L
283 K
This answer supports our expectation from Charles’s law, namely, that raising the gas temperature (from 283 K to 303 K) at a
constant pressure will yield an increase in its volume (from 0.300 L to 0.321 L).
Exercise 7.2.2
A sample of oxygen, O2, occupies 32.2 mL at 30 °C and 452 torr. What volume will it occupy at –70 °C and the same pressure?
Answer
7.2.4 https://chem.libretexts.org/@go/page/217290
21.6 mL
3
131.7 cm × 273.15 K
Rearrangement gives T 2 =
3
= 239.8 K
150.0 cm
Subtracting 273.15 from 239.8 K, we find that the temperature of the boiling ammonia on the Celsius scale is –33.4 °C.
Exercise 7.2.3
What is the volume of a sample of ethane at 467 K and 1.1 atm if it occupies 405 mL at 298 K and 1.1 atm?
Answer
635 mL
7.2.5 https://chem.libretexts.org/@go/page/217290
Figure 7.2.6 : When a gas occupies a smaller volume, it exerts a higher pressure; when it occupies a larger volume, it exerts a
lower pressure (assuming the amount of gas and the temperature do not change). Since P and V are inversely proportional, a graph
1
of vs. V is linear.
P
Unlike the P-T and V-T relationships, pressure and volume are not directly proportional to each other. Instead, P and V exhibit
inverse proportionality: Increasing the pressure results in a decrease of the volume of the gas. Mathematically this can be written:
with k being a constant. Graphically, this relationship is shown by the straight line that results when plotting the inverse of the
1 1
pressure ( ) versus the volume (V), or the inverse of volume ( ) versus the pressure (P). Graphs with curved lines are
P V
difficult to read accurately at low or high values of the variables, and they are more difficult to use in fitting theoretical equations
and parameters to experimental data. For those reasons, scientists often try to find a way to “linearize” their data. If we plot P
versus V, we obtain a hyperbola (Figure 7.2.7).
7.2.6 https://chem.libretexts.org/@go/page/217290
Figure 7.2.7 : The relationship between pressure and volume is inversely proportional. (a) The graph of P vs. V is a hyperbola,
1
whereas (b) the graph of ( ) vs. V is linear.
P
The relationship between the volume and pressure of a given amount of gas at constant temperature was first published by the
English natural philosopher Robert Boyle over 300 years ago. It is summarized in the statement now known as Boyle’s law: The
volume of a given amount of gas held at constant temperature is inversely proportional to the pressure under which it is measured.
Figure 7.2.7 : An illustration of Boyle's Law, showing that as volume decreases, pressure increases, and vice versa. This file is in
the public domain because it was created by NASA . NASA's copyright policy states that "NASA material is not
copyrighted unless otherwise specified ". ( NASA copyright policy and JPL Image Use Policy ).
7.2.7 https://chem.libretexts.org/@go/page/217290
P1 V1 = P2 V2 or 13.0 psi × 15.0 mL = P2 × 7.5 mL (7.2.5)
Solving:
It was more difficult to estimate well from the P-V graph, so (a) is likely more inaccurate than (b) or (c). The calculation will be
as accurate as the equation and measurements allow.
Exercise 7.2.4
The sample of gas has a volume of 30.0 mL at a pressure of 6.5 psi. Determine the volume of the gas at a pressure of 11.0 psi,
using:
a. the P-V graph in Figure 7.2.6a
1
b. the vs. V graph in Figure 7.2.6b
P
c. the Boyle’s law equation
Comment on the likely accuracy of each method.
Answer a
about 17–18 mL
Answer b
~18 mL
Answer c
17.7 mL; it was more difficult to estimate well from the P-V graph, so (a) is likely more inaccurate than (b); the calculation
will be as accurate as the equation and measurements allow
The ideal gas law is easy to remember and apply in solving problems, as long as you get
the proper values and units for the gas constant, R.
Chemistry in Everyday Life: Breathing and Boyle’s Law
What do you do about 20 times per minute for your whole life, without break, and often without even being aware of it? The
answer, of course, is respiration, or breathing. How does it work? It turns out that the gas laws apply here. Your lungs take in gas
that your body needs (oxygen) and get rid of waste gas (carbon dioxide). Lungs are made of spongy, stretchy tissue that expands
and contracts while you breathe. When you inhale, your diaphragm and intercostal muscles (the muscles between your ribs)
contract, expanding your chest cavity and making your lung volume larger. The increase in volume leads to a decrease in
pressure (Boyle’s law). This causes air to flow into the lungs (from high pressure to low pressure). When you exhale, the
process reverses: Your diaphragm and rib muscles relax, your chest cavity contracts, and your lung volume decreases, causing
the pressure to increase (Boyle’s law again), and air flows out of the lungs (from high pressure to low pressure). You then
breathe in and out again, and again, repeating this Boyle’s law cycle for the rest of your life (Figure 7.2.8).
7.2.8 https://chem.libretexts.org/@go/page/217290
Figure 7.2.8 : Breathing occurs because expanding and contracting lung volume creates small pressure differences between
your lungs and your surroundings, causing air to be drawn into and forced out of your lungs.
Mathematical relationships can also be determined for the other variable pairs, such as P versus n, and n versus T.
Combining these four laws yields the ideal gas law, a relation between the pressure, volume, temperature, and number of moles of
a gas:
7.2.9 https://chem.libretexts.org/@go/page/217290
P V = nRT (7.2.8)
where P is the pressure of a gas, V is its volume, n is the number of moles of the gas, T is its temperature on the kelvin scale, and R
is a constant called the ideal gas constant or the universal gas constant. The units used to express pressure, volume, and temperature
will determine the proper form of the gas constant as required by dimensional analysis, the most commonly encountered values
being 0.08206 L atm mol–1 K–1 and 8.314 kPa L mol–1 K–1.
Gases whose properties of P, V, and T are accurately described by the ideal gas law (or the other gas laws) are said to exhibit ideal
behavior or to approximate the traits of an ideal gas. An ideal gas is a hypothetical construct that may be used along with kinetic
molecular theory to effectively explain the gas laws as will be described in a later module of this chapter. Although all the
calculations presented in this module assume ideal behavior, this assumption is only reasonable for gases under conditions of
relatively low pressure and high temperature. In the final module of this chapter, a modified gas law will be introduced that
accounts for the non-ideal behavior observed for many gases at relatively high pressures and low temperatures.
The ideal gas equation contains five terms, the gas constant R and the variable properties P, V, n, and T. Specifying any four of
these terms will permit use of the ideal gas law to calculate the fifth term as demonstrated in the following example exercises.
If we choose to use R = 0.08206 L atm mol–1 K–1, then the amount must be in moles, temperature must be in kelvin, and
pressure must be in atm.
Converting into the “right” units:
1 mol
n = 655 g CH4 × = 40.8 mol (7.2.9)
16.043 g CH4
1atm
P = 745 torr × = 0.980 atm (7.2.11)
760 torr
–1 –1
nRT (40.8 mol )(0.08206 L atmmol K )(298 K)
3
V = = = 1.02 × 10 L (7.2.12)
P 0.980 atm
It would require 1020 L (269 gal) of gaseous methane at about 1 atm of pressure to replace 1 gal of gasoline. It requires a large
container to hold enough methane at 1 atm to replace several gallons of gasoline.
Exercise 7.2.5
Calculate the pressure in bar of 2520 moles of hydrogen gas stored at 27 °C in the 180-L storage tank of a modern hydrogen-
powered car.
Answer
350 bar
If the number of moles of an ideal gas are kept constant under two different sets of conditions, a useful mathematical relationship
P1 V1 P2 V2
called the combined gas law is obtained: = using units of atm, L, and K. Both sets of conditions are equal to the
T1 T2
product of n × R (where n = the number of moles of the gas and R is the ideal gas law constant).
7.2.10 https://chem.libretexts.org/@go/page/217290
When filled with air, a typical scuba tank with a volume of 13.2 L has a pressure of 153 atm (Figure 7.2.9). If the water
temperature is 27 °C, how many liters of air will such a tank provide to a diver’s lungs at a depth of approximately 70 feet in the
ocean where the pressure is 3.13 atm?
Figure 7.2.9 : Scuba divers use compressed air to breathe while underwater. (credit: modification of work by Mark Goodchild)
Letting 1 represent the air in the scuba tank and 2 represent the air in the lungs, and noting that body temperature (the
temperature the air will be in the lungs) is 37 °C, we have:
(Note: Be advised that this particular example is one in which the assumption of ideal gas behavior is not very reasonable, since
it involves gases at relatively high pressures and low temperatures. Despite this limitation, the calculated volume can be viewed
as a good “ballpark” estimate.)
Exercise 7.2.6
A sample of ammonia is found to occupy 0.250 L under laboratory conditions of 27 °C and 0.850 atm. Find the volume of this
sample at 0 °C and 1.00 atm.
Answer
0.193 L
chemistry in everyday life: The Interdependence between Ocean Depth and Pressure in Scuba
Diving
Whether scuba diving at the Great Barrier Reef in Australia (shown in Figure 7.2.10) or in the Caribbean, divers must
understand how pressure affects a number of issues related to their comfort and safety.
7.2.11 https://chem.libretexts.org/@go/page/217290
Figure 7.2.10 : Scuba divers, whether at the Great Barrier Reef or in the Caribbean, must be aware of buoyancy, pressure
equalization, and the amount of time they spend underwater, to avoid the risks associated with pressurized gases in the body.
(credit: Kyle Taylor).
Pressure increases with ocean depth, and the pressure changes most rapidly as divers reach the surface. The pressure a diver
experiences is the sum of all pressures above the diver (from the water and the air). Most pressure measurements are given in
units of atmospheres, expressed as “atmospheres absolute” or ATA in the diving community: Every 33 feet of salt water
represents 1 ATA of pressure in addition to 1 ATA of pressure from the atmosphere at sea level. As a diver descends, the
increase in pressure causes the body’s air pockets in the ears and lungs to compress; on the ascent, the decrease in pressure
causes these air pockets to expand, potentially rupturing eardrums or bursting the lungs. Divers must therefore undergo
equalization by adding air to body airspaces on the descent by breathing normally and adding air to the mask by breathing out of
the nose or adding air to the ears and sinuses by equalization techniques; the corollary is also true on ascent, divers must release
air from the body to maintain equalization. Buoyancy, or the ability to control whether a diver sinks or floats, is controlled by
the buoyancy compensator (BCD). If a diver is ascending, the air in his BCD expands because of lower pressure according to
Boyle’s law (decreasing the pressure of gases increases the volume). The expanding air increases the buoyancy of the diver, and
she or he begins to ascend. The diver must vent air from the BCD or risk an uncontrolled ascent that could rupture the lungs. In
descending, the increased pressure causes the air in the BCD to compress and the diver sinks much more quickly; the diver must
add air to the BCD or risk an uncontrolled descent, facing much higher pressures near the ocean floor. The pressure also impacts
how long a diver can stay underwater before ascending. The deeper a diver dives, the more compressed the air that is breathed
because of increased pressure: If a diver dives 33 feet, the pressure is 2 ATA and the air would be compressed to one-half of its
original volume. The diver uses up available air twice as fast as at the surface.
7.2.12 https://chem.libretexts.org/@go/page/217290
Figure 7.2.11 : Since the number of moles in a given volume of gas varies with pressure and temperature changes, chemists use
standard temperature and pressure (273.15 K and 1 atm or 101.325 kPa) to report properties of gases.
Summary
Video 7.2.2 : An overview of the ideal gas law and STP and a demonstration of the ideal gas law.
The behavior of gases can be described by several laws based on experimental observations of their properties. The pressure of a
given amount of gas is directly proportional to its absolute temperature, provided that the volume does not change (Amontons’s
law). The volume of a given gas sample is directly proportional to its absolute temperature at constant pressure (Charles’s law). The
volume of a given amount of gas is inversely proportional to its pressure when temperature is held constant (Boyle’s law). Under
the same conditions of temperature and pressure, equal volumes of all gases contain the same number of molecules (Avogadro’s
law).
The equations describing these laws are special cases of the ideal gas law, PV = nRT, where P is the pressure of the gas, V is its
volume, n is the number of moles of the gas, T is its kelvin temperature, and R is the ideal (universal) gas constant.
7.2.13 https://chem.libretexts.org/@go/page/217290
Key Equations
PV = nRT
Glossary
absolute zero
temperature at which the volume of a gas would be zero according to Charles’s law.
Amontons’s law
(also, Gay-Lussac’s law) pressure of a given number of moles of gas is directly proportional to its kelvin temperature when the
volume is held constant
Avogadro’s law
volume of a gas at constant temperature and pressure is proportional to the number of gas molecules
Boyle’s law
volume of a given number of moles of gas held at constant temperature is inversely proportional to the pressure under which it
is measured
Charles’s law
volume of a given number of moles of gas is directly proportional to its kelvin temperature when the pressure is held constant
ideal gas
hypothetical gas whose physical properties are perfectly described by the gas laws
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
7.2: The Gas Laws is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
7.2.14 https://chem.libretexts.org/@go/page/217290
7.2.1: Practice Problems- The Gas Laws
PROBLEM 7.2.1.1
Sometimes leaving a bicycle in the sun on a hot day will cause a blowout. Why?
Answer
As temperature of a gas increases, pressure will also increase based on the ideal gas law. The volume of the tire can only
expand so much before the rubber gives and releases the build up of pressure.
PROBLEM 7.2.1.2
Explain how the volume of the bubbles exhausted by a scuba diver change as they rise to the surface, assuming that they remain
intact.
Answer
As the bubbles rise, the pressure decreases, so their volume increases as suggested by Boyle’s law.
PROBLEM 7.2.1.3
One way to state Boyle’s law is “All other things being equal, the pressure of a gas is inversely proportional to its volume.”
(a) What is the meaning of the term “inversely proportional?”
(b) What are the “other things” that must be equal?
Answer a
The pressure of the gas increases as the volume decreases
Answer b
amount of gas, temperature
PROBLEM 7.2.1.4
An alternate way to state Avogadro’s law is “All other things being equal, the number of molecules in a gas is directly
proportional to the volume of the gas.”
a. What is the meaning of the term “directly proportional?”
b. What are the “other things” that must be equal?
Answer a
The number of particles in the gas increases as the volume increases
Answer b
temperature, pressure
PROBLEM 7.2.1.5
A spray can is used until it is empty except for the propellant gas, which has a pressure of 1344 torr at 23 °C. If the can is
thrown into a fire (T = 475 °C), what will be the pressure in the hot can?
Answer
3.40 × 103 torr
7.2.1.1 https://chem.libretexts.org/@go/page/217291
Problem 7.2.5
PROBLEM 7.2.1.6
What is the temperature of an 11.2-L sample of carbon monoxide, CO, at 744 torr if it occupies 13.3 L at 55 °C and 744 torr?
Answer
276°K
3°C
PROBLEM 7.2.1.7
A 2.50-L volume of hydrogen measured at –196 °C is warmed to 100 °C. Calculate the volume of the gas at the higher
temperature, assuming no change in pressure.
Answer
12.1 L
Problem 7.2.7
PROBLEM 7.2.1.8
7.2.1.2 https://chem.libretexts.org/@go/page/217291
A balloon inflated with three breaths of air has a volume of 1.7 L. At the same temperature and pressure, what is the volume of
the balloon if five more same-sized breaths are added to the balloon?
Answer
4.5 L
PROBLEM 7.2.1.9
A weather balloon contains 8.80 moles of helium at a pressure of 0.992 atm and a temperature of 25 °C at ground level. What is
the volume of the balloon under these conditions?
Answer
217 L
Problem 7.2.9
PROBLEM 7.2.1.10
How many grams of gas are present in each of the following cases?
a. 0.100 L of CO2 at 307 torr and 26 °C
b. 8.75 L of C2H4, at 378.3 kPa and 483 K
c. 221 mL of Ar at 0.23 torr and –54 °C
Answer a
7.24 × 10–2 g
Answer b
23.1 g
Answer c
1.5 × 10–4 g
PROBLEM 7.2.1.11
A high altitude balloon is filled with 1.41 × 104 L of hydrogen at a temperature of 21 °C and a pressure of 745 torr. What is the
volume of the balloon at a height of 20 km, where the temperature is –48 °C and the pressure is 63.1 torr?
Answer
7.2.1.3 https://chem.libretexts.org/@go/page/217291
1.2741 × 105 L or more correctly to 3 significant figures 1.27 × 105 L
Problem 7.2.11
PROBLEM 7.2.1.12
While resting, the average 70-kg human male consumes 14 L of pure O2 per hour at 25 °C and 100 kPa. How many moles of O2
are consumed by a 70 kg man while resting for 1.0 h?
Answer
0.0565 mol
PROBLEM 7.2.1.13
A balloon that is 100.21 L at 21 °C and 0.981 atm is released and just barely clears the top of Mount Crumpet in British
Columbia. If the final volume of the balloon is 144.53 L at a temperature of 5.24 °C, what is the pressure experienced by the
balloon as it clears Mount Crumpet?
Answer
0.644 atm
Problem 7.2.13
PROBLEM 7.2.1.14
7.2.1.4 https://chem.libretexts.org/@go/page/217291
If the temperature of a fixed amount of a gas is doubled at constant volume, what happens to the pressure?
Answer
Temperature and Pressure are directly proportional. Pressure will also have to increase (doubling).
PROBLEM 7.2.1.15
If the volume of a fixed amount of a gas is tripled at constant temperature, what happens to the pressure?
Answer
Volume and pressure are inversely proportional. The pressure decreases by a factor of 3.
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
7.2.1: Practice Problems- The Gas Laws is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
7.2.1.5 https://chem.libretexts.org/@go/page/217291
7.3: Applications of the Ideal Gas Law and Partial Pressures
Skills to Develop
Use the ideal gas law to compute gas densities and molar masses
Perform stoichiometric calculations involving gaseous substances
State Dalton’s law of partial pressures and use it in calculations involving gaseous mixtures
The study of the chemical behavior of gases was part of the basis of perhaps the most fundamental chemical revolution in history.
French nobleman Antoine Lavoisier, widely regarded as the “father of modern chemistry,” changed chemistry from a qualitative to
a quantitative science through his work with gases. He discovered the law of conservation of matter, discovered the role of oxygen
in combustion reactions, determined the composition of air, explained respiration in terms of chemical reactions, and more. He was
a casualty of the French Revolution, guillotined in 1794. Of his death, mathematician and astronomer Joseph-Louis Lagrange said,
“It took the mob only a moment to remove his head; a century will not suffice to reproduce it."
As described in an earlier chapter of this text, we can turn to chemical stoichiometry for answers to many of the questions that ask
“How much?” We can answer the question with masses of substances or volumes of solutions. However, we can also answer this
question another way: with volumes of gases. We can use the ideal gas equation to relate the pressure, volume, temperature, and
number of moles of a gas. Here we will combine the ideal gas equation with other equations to find gas density and molar mass.
We will deal with mixtures of different gases, and calculate amounts of substances in reactions involving gases. This section will
not introduce any new material or ideas, but will provide examples of applications and ways to integrate concepts we have already
discussed.
Density of a Gas
m
Recall that the density of a gas is its mass to volume ratio, ρ = . Therefore, if we can determine the mass of some volume of a
V
gas, we will get its density. The density of an unknown gas can used to determine its molar mass and thereby assist in its
identification. The ideal gas law, PV = nRT, provides us with a means of deriving such a mathematical formula to relate the density
of a gas to its volume in the proof shown in Example 7.3.1.
Example 7.3.1 : Derivation of a Density Formula from the Ideal Gas Law
Use PV = nRT to derive a formula for the density of gas in g/L
Solution
P V = nRT (7.3.1)
Multiply each side of the equation by the molar mass, ℳ. When moles are multiplied by ℳ in g/mol, g are obtained:
n P
(M) ( ) =( ) (M) (7.3.3)
V RT
PM
M/V = ρ = (7.3.4)
RT
Exercise 7.3.1
A gas was found to have a density of 0.0847 g/L at 17.0 °C and a pressure of 760 torr. What is its molar mass? What is the gas?
Answer
PM
ρ =
RT
7.3.1 https://chem.libretexts.org/@go/page/217292
1 atm M
0.0847 g/L = 760 torr × × × 290 K (7.3.5)
760 torr 0.0821 L atm /mol K
ℳ = 2.02 g/mol; therefore, the gas must be hydrogen (H2, 2.02 g/mol)
We must specify both the temperature and the pressure of a gas when calculating its density because the number of moles of a gas
(and thus the mass of the gas) in a liter changes with temperature or pressure. Gas densities are often reported at STP.
Example 7.3.2 : Empirical/Molecular Formula Problems Using the Ideal Gas Law and Density of a
Gas
Cyclopropane, a gas once used with oxygen as a general anesthetic, is composed of 85.7% carbon and 14.3% hydrogen by mass.
Find the empirical formula. If 1.56 g of cyclopropane occupies a volume of 1.00 L at 0.984 atm and 50 °C, what is the
molecular formula for cyclopropane?
Solution
Strategy:
First solve the empirical formula problem using methods discussed earlier. Assume 100 g and convert the percentage of each
element into grams. Determine the number of moles of carbon and hydrogen in the 100-g sample of cyclopropane. Divide by the
smallest number of moles to relate the number of moles of carbon to the number of moles of hydrogen. In the last step, realize
that the smallest whole number ratio is the empirical formula:
1 mol C 7.136
85.7 g C × = 7.136 mol C = 1.00 mol C (7.3.6)
12.01 g C 7.136
1 mol H 14.158
14.3 g H × = 14.158 mol H = 1.98 mol H (7.3.7)
1.01 g H 7.136
M 42.0
ℳ = 42.0 g/mol, = = 2.99 , so (3)(CH2) = C3H6 (molecular formula)
EM 14.03
Exercise 7.3.2
Acetylene, a fuel used welding torches, is comprised of 92.3% C and 7.7% H by mass. Find the empirical formula. If 1.10 g of
acetylene occupies of volume of 1.00 L at 1.15 atm and 59.5 °C, what is the molecular formula for acetylene?
Answer
Empirical formula, CH; Molecular formula, C2H2
7.3.2 https://chem.libretexts.org/@go/page/217292
and then combined with the molar mass equation to yield:
mRT
M= (7.3.11)
PV
This equation can be used to derive the molar mass of a gas from measurements of its pressure, volume, temperature, and mass.
Figure 7.3.1 : When the volatile liquid in the flask is heated past its boiling point, it becomes gas and drives air out of the flask.
At t , the flask is filled with volatile liquid gas at the same pressure as the atmosphere. If the flask is then cooled to room
l⟶g
temperature, the gas condenses and the mass of the gas that filled the flask, and is now liquid, can be measured. (credit:
modification of work by Mark Ott)
Using this procedure, a sample of chloroform gas weighing 0.494 g is collected in a flask with a volume of 129 cm3 at 99.6 °C
when the atmospheric pressure is 742.1 mm Hg. What is the approximate molar mass of chloroform?
Solution
Since
m
M=
n
and
PV
n =
RT
then
mRT (0.494 g) × 0.08206 L ⋅ atm/mol K × 372.8 K
M= = = 120 g/mol (7.3.13)
PV 0.976 atm × 0.129 L
Exercise 7.3.3
A sample of phosphorus that weighs 3.243 × 10−2 g exerts a pressure of 31.89 kPa in a 56.0-mL bulb at 550 °C. What are the
molar mass and molecular formula of phosphorus vapor?
Answer
7.3.3 https://chem.libretexts.org/@go/page/217292
124 g/mol P4
PT otal = PA + PB + PC +. . . = ∑ Pi (7.3.14)
In the equation PTotal is the total pressure of a mixture of gases, PA is the partial pressure of gas A; PB is the partial pressure of gas
B; PC is the partial pressure of gas C; and so on.
Figure 7.3.2 : If equal-volume cylinders containing gas A at a pressure of 300 kPa, gas B at a pressure of 600 kPa, and gas C at a
pressure of 450 kPa are all combined in the same-size cylinder, the total pressure of the mixture is 1350 kPa.
The partial pressure of gas A is related to the total pressure of the gas mixture via its mole fraction (X), a unit of concentration
defined as the number of moles of a component of a solution divided by the total number of moles of all components:
nA
PA = XA × PT otal where XA = (7.3.15)
nT otal
where PA, XA, and nA are the partial pressure, mole fraction, and number of moles of gas A, respectively, and nTotal is the number of
moles of all components in the mixture.
−3 −1 −1
(2.50 × 10 mol)(0.08206 L atm mol K )(308 K )
−3
PH = = 6.32 × 10 atm (7.3.16)
2
10.0 L
−3 −1 −1
(1.00 × 10 mol )(0.08206 L atm mol K )(308 K )
−3
PHe = = 2.53 × 10 atm (7.3.17)
10.0 L
7.3.4 https://chem.libretexts.org/@go/page/217292
−4 −1 −1
(3.00 × 10 mol )(0.08206 L atm mol K )(308 K )
−4
PNe = = 7.58 × 10 atm (7.3.18)
10.0 L
Exercise 7.3.4
A 5.73-L flask at 25 °C contains 0.0388 mol of N2, 0.147 mol of CO, and 0.0803 mol of H2. What is the total pressure in the
flask in atmospheres?
Answer
1.137 atm
Here is another example of this concept, but dealing with mole fraction calculations.
For O2,
nO 2.83mol
2
XO = = = 0.252
2
nT otal (2.83 + 8.41) mol
and
For N2O,
nN O 8.41 mol
2
XN2 O = = = 0.748
nT otal (2.83 + 8.41) mol
and
Exercise 7.3.5
What is the pressure of a mixture of 0.200 g of H2, 1.00 g of N2, and 0.820 g of Ar in a container with a volume of 2.00 L at 20
°C?
Answer
1.87 atm
7.3.5 https://chem.libretexts.org/@go/page/217292
Collection of Gases over Water
A simple way to collect gases that do not react with water is to capture them in a bottle that has been filled with water and inverted
into a dish filled with water. The pressure of the gas inside the bottle can be made equal to the air pressure outside by raising or
lowering the bottle. When the water level is the same both inside and outside the bottle (Figure 7.3.3), the pressure of the gas is
equal to the atmospheric pressure, which can be measured with a barometer.
Figure 7.3.3 : When a reaction produces a gas that is collected above water, the trapped gas is a mixture of the gas produced by
the reaction and water vapor. If the collection flask is appropriately positioned to equalize the water levels both within and outside
the flask, the pressure of the trapped gas mixture will equal the atmospheric pressure outside the flask (see the earlier discussion of
manometers).
However, there is another factor we must consider when we measure the pressure of the gas by this method. Water evaporates and
there is always gaseous water (water vapor) above a sample of liquid water. As a gas is collected over water, it becomes saturated
with water vapor and the total pressure of the mixture equals the partial pressure of the gas plus the partial pressure of the water
vapor. The pressure of the pure gas is therefore equal to the total pressure minus the pressure of the water vapor—this is referred to
as the “dry” gas pressure, that is, the pressure of the gas only, without water vapor.
Figure 7.3.4 : This graph shows the vapor pressure of water at sea level as a function of temperature.
The vapor pressure of water, which is the pressure exerted by water vapor in equilibrium with liquid water in a closed container,
depends on the temperature (Figure 7.3.4); more detailed information on the temperature dependence of water vapor can be found
in Table 7.3.1, and vapor pressure will be discussed in more detail in the next chapter on liquids.
Table 7.3.1: Vapor Pressure of Ice and Water in Various Temperatures at Sea Level
Temperature Temperature Temperature
Pressure (torr) Pressure (torr) Pressure (torr)
(°C) (°C) (°C)
7.3.6 https://chem.libretexts.org/@go/page/217292
Temperature Temperature Temperature
Pressure (torr) Pressure (torr) Pressure (torr)
(°C) (°C) (°C)
The pressure of water vapor above a sample of liquid water at 26 °C is 25.2 torr (Appendix E), so:
Exercise 7.3.6
A sample of oxygen collected over water at a temperature of 29.0 °C and a pressure of 764 torr has a volume of 0.560 L. What
volume would the dry oxygen have under the same conditions of temperature and pressure?
Answer
0.583 L
7.3.7 https://chem.libretexts.org/@go/page/217292
Avogadro’s Law Revisited
Sometimes we can take advantage of a simplifying feature of the stoichiometry of gases that solids and solutions do not exhibit: All
gases that show ideal behavior contain the same number of molecules in the same volume (at the same temperature and pressure).
Thus, the ratios of volumes of gases involved in a chemical reaction are given by the coefficients in the equation for the reaction,
provided that the gas volumes are measured at the same temperature and pressure.
We can extend Avogadro’s law (that the volume of a gas is directly proportional to the number of moles of the gas) to chemical
reactions with gases: Gases combine, or react, in definite and simple proportions by volume, provided that all gas volumes are
measured at the same temperature and pressure. For example, since nitrogen and hydrogen gases react to produce ammonia gas
according to
N (g) + 3 H (g) ⟶ 2 NH (g) (7.3.23)
2 2 3
a given volume of nitrogen gas reacts with three times that volume of hydrogen gas to produce two times that volume of ammonia
gas, if pressure and temperature remain constant.
The explanation for this is illustrated in Figure 7.3.4. According to Avogadro’s law, equal volumes of gaseous N2, H2, and NH3, at
the same temperature and pressure, contain the same number of molecules. Because one molecule of N2 reacts with three
molecules of H2 to produce two molecules of NH3, the volume of H2 required is three times the volume of N2, and the volume of
NH3 produced is two times the volume of N2.
Figure 7.3.5 : One volume of N2 combines with three volumes of H2 to form two volumes of NH3.
From the equation, we see that one volume of C3H8 will react with five volumes of O2:
5 LO
2
2.7 L C3 H8 × = 13.5 L O (7.3.26)
2
1 L C3 H8
7.3.8 https://chem.libretexts.org/@go/page/217292
Exercise 7.3.7
An acetylene tank for an oxyacetylene welding torch provides 9340 L of acetylene gas, C2H2, at 0 °C and 1 atm. How many
tanks of oxygen, each providing 7.00 × 103 L of O2 at 0 °C and 1 atm, will be required to burn the acetylene?
2C H +5 O ⟶ 4 CO +2 H O
2 2 2 2 2
Answer
3.34 tanks (2.34 × 104 L)
Solution
Because equal volumes of H2 and NH3 contain equal numbers of molecules and each three molecules of H2 that react produce
two molecules of NH3, the ratio of the volumes of H2 and NH3 will be equal to 3:2. Two volumes of NH3, in this case in units of
billion ft3, will be formed from three volumes of H2:
3
3 billion f t H2
3 3 3
683 billion f t NH3 × = 1.02 × 10 billion f t H2 (7.3.27)
3
2 billion f t NH3
The manufacture of 683 billion ft3 of NH3 required 1020 billion ft3 of H2. (At 25 °C and 1 atm, this is the volume of a cube with
an edge length of approximately 1.9 miles.)
Exercise 7.3.8
What volume of O2(g) measured at 25 °C and 760 torr is required to react with 17.0 L of ethylene, C2H4(g), measured under the
same conditions of temperature and pressure? The products are CO2 and water vapor.
Answer
51.0 L
Solution
To convert from the mass of gallium to the volume of H2(g), we need to do something like this:
1 mol Ga 3 mol H2
8.88 g Ga × × = 0.191 mol H2 (7.3.29)
69.723 g Ga 2 mol Ga
7.3.9 https://chem.libretexts.org/@go/page/217292
−1 −1
nRT 0.191 mol × 0.08206 L atm mol K × 300 K
VH2 = ( ) = = 4.94 L (7.3.30)
P 0.951 atm
H2
Exercise 7.3.9
Sulfur dioxide is an intermediate in the preparation of sulfuric acid. What volume of SO2 at 343 °C and 1.21 atm is produced by
burning l.00 kg of sulfur in oxygen?
Answer
1.30 × 103 L
Figure 7.3.6 : Greenhouse gases trap enough of the sun’s energy to make the planet habitable—this is known as the greenhouse
effect. Human activities are increasing greenhouse gas levels, warming the planet and causing more extreme weather events.
There is strong evidence from multiple sources that higher atmospheric levels of CO2 are caused by human activity, with fossil
3
fuel burning accounting for about of the recent increase in CO2. Reliable data from ice cores reveals that CO2 concentration
4
in the atmosphere is at the highest level in the past 800,000 years; other evidence indicates that it may be at its highest level in
20 million years. In recent years, the CO2 concentration has increased from historical levels of below 300 ppm to almost 400
ppm today (Figure 7.3.7).
7.3.10 https://chem.libretexts.org/@go/page/217292
Figure 7.3.7 : CO2 levels over the past 700,000 years were typically from 200–300 ppm, with a steep, unprecedented increase
over the past 50 years.
Figure 7.3.8 : Susan Solomon’s research focuses on climate change and has been instrumental in determining the cause of the
ozone hole over Antarctica. (credit: National Oceanic and Atmospheric Administration)
Summary
7.3.11 https://chem.libretexts.org/@go/page/217292
Partial Pressures & Vapor Pressure: Cra…
Cra…
Video 7.3.1 : An overview of the topics of partial pressure, vapor pressure, and collecting gases over water.
The ideal gas law can be used to derive a number of convenient equations relating directly measured quantities to properties of
interest for gaseous substances and mixtures. Appropriate rearrangement of the ideal gas equation may be made to permit the
calculation of gas densities and molar masses. Dalton’s law of partial pressures may be used to relate measured gas pressures for
gaseous mixtures to their compositions. Avogadro’s law may be used in stoichiometric computations for chemical reactions
involving gaseous reactants or products.
Key Equations
PTotal = PA + PB + PC + … = ƩiPi
PA = XA PTotal
nA
XA =
nT otal
Footnotes
1. “Quotations by Joseph-Louis Lagrange,” last modified February 2006, accessed February 10, 2015, http://www-history.mcs.st-
andrews.ac.../Lagrange.html
Glossary
partial pressure
pressure exerted by an individual gas in a mixture
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
7.3.12 https://chem.libretexts.org/@go/page/217292
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
7.3: Applications of the Ideal Gas Law and Partial Pressures is shared under a not declared license and was authored, remixed, and/or curated by
LibreTexts.
7.3.13 https://chem.libretexts.org/@go/page/217292
7.3.1: Practice Problems- Applications of the Ideal Gas Law
PROBLEM 7.3.1.1
What is the density of laughing gas, dinitrogen monoxide, N2O, at a temperature of 325 K and a pressure of 113.0 kPa?
Answer
1.84 g/L
PROBLEM 7.3.1.2
Calculate the density of Freon 12, CF2Cl2, at 30.0 °C and 0.954 atm.
Answer
4.64 g/L
Problem 7.3.2
PROBLEM 7.3.1.3
Which is denser at the same temperature and pressure, dry air or air saturated with water vapor? Explain.
Answer
air saturated with water vapor - it has a higher molar mass
PROBLEM 7.3.1.4
A cylinder of O2(g) used in breathing by emphysema patients has a volume of 3.00 L at a pressure of 10.0 atm. If the
temperature of the cylinder is 28.0 °C, what mass of oxygen is in the cylinder?
Answer
38.8 g
7.3.1.1 https://chem.libretexts.org/@go/page/217293
Problem 7.3.4
PROBLEM 7.3.1.5
What is the molar mass of a gas if 0.0494 g of the gas occupies a volume of 0.100 L at a temperature 26 °C and a pressure of
307 torr?
Answer
30.0 g/mol
PROBLEM 7.3.1.6
What is the molar mass of a gas if 0.281 g of the gas occupies a volume of 125 mL at a temperature 126 °C and a pressure of
777 torr?
Answer
72.0 g/mol
Problem 7.3.6
PROBLEM 7.3.1.7
The density of a certain gaseous fluoride of phosphorus is 3.93 g/L at STP. Calculate the molar mass of this fluoride and
determine its molecular formula.
7.3.1.2 https://chem.libretexts.org/@go/page/217293
Answer
88.1 g mol−1; PF3
PROBLEM 7.3.1.8
What is the molecular formula of a compound that contains 39% C, 45% N, and 16% H if 0.157 g of the compound occupies
125 mL with a pressure of 99.5 kPa at 22 °C?
Answer
H5CN
Problem 7.3.8
PROBLEM 7.3.1.9
A sample of gas isolated from unrefined petroleum contains 90.0% CH4, 8.9% C2H6, and 1.1% C3H8 at a total pressure of 307.2
kPa. What is the partial pressure of each component of this gas? (The percentages given indicate the percent of the total pressure
that is due to each component.)
Answer
CH4: 276 kPa; C2H6: 27 kPa; C3H8: 3.4 kPa
PROBLEM 7.3.1.10
Automobile air bags are inflated with nitrogen gas, which is formed by the decomposition of solid sodium azide (NaN3). The
other product is sodium metal. Calculate the volume of nitrogen gas at 27 °C and 756 torr formed by the decomposition of 125 g
of sodium azide.
Answer
71.26 L
7.3.1.3 https://chem.libretexts.org/@go/page/217293
Problem 7.3.10
PROBLEM 7.3.1.11
A sample of a compound of xenon and fluorine was confined in a bulb with a pressure of 18 torr. Hydrogen was added to the
bulb until the pressure was 72 torr. Passage of an electric spark through the mixture produced Xe and HF. After the HF was
removed by reaction with solid KOH, the final pressure of xenon and unreacted hydrogen in the bulb was 36 torr. What is the
empirical formula of the xenon fluoride in the original sample? (Note: Xenon fluorides contain only one xenon atom per
molecule.)
Answer
XeF2
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
7.3.1: Practice Problems- Applications of the Ideal Gas Law is shared under a not declared license and was authored, remixed, and/or curated by
LibreTexts.
7.3.1.4 https://chem.libretexts.org/@go/page/217293
7.4: Kinetic Molecular Theory
Skills to Develop
State the postulates of the kinetic-molecular theory
Use this theory’s postulates to explain the gas laws
The gas laws that we have seen to this point, as well as the ideal gas equation, are empirical, that is, they have been derived from experimental
observations. The mathematical forms of these laws closely describe the macroscopic behavior of most gases at pressures less than about 1 or 2 atm.
Although the gas laws describe relationships that have been verified by many experiments, they do not tell us why gases follow these relationships.
The kinetic molecular theory (KMT) is a simple microscopic model that effectively explains the gas laws described in previous modules of this chapter.
This theory is based on the following five postulates described here. (Note: The term “molecule” will be used to refer to the individual chemical species
that compose the gas, although some gases are composed of atomic species, for example, the noble gases.)
1. Gases are composed of molecules that are in continuous motion, travelling in straight lines and changing direction only when they collide with other
molecules or with the walls of a container.
2. The molecules composing the gas are negligibly small compared to the distances between them.
3. The pressure exerted by a gas in a container results from collisions between the gas molecules and the container walls.
4. Gas molecules exert no attractive or repulsive forces on each other or the container walls; therefore, their collisions are elastic (do not involve a loss of
energy).
5. The average kinetic energy of the gas molecules is proportional to the kelvin temperature of the gas.
The test of the KMT and its postulates is its ability to explain and describe the behavior of a gas. The various gas laws can be derived from the
assumptions of the KMT, which have led chemists to believe that the assumptions of the theory accurately represent the properties of gas molecules. We
will first look at the individual gas laws (Boyle’s, Charles’s, Amontons’s, Avogadro’s, and Dalton’s laws) conceptually to see how the KMT explains
them. Then, we will more carefully consider the relationships between molecular masses, speeds, and kinetic energies with temperature, and explain
Graham’s law.
7.4.1 https://chem.libretexts.org/@go/page/217294
Figure 7.4.1 : (a) When gas temperature increases, gas pressure increases due to increased force and frequency of molecular collisions. (b) When volume
decreases, gas pressure increases due to increased frequency of molecular collisions. (c) When the amount of gas increases at a constant pressure, volume
increases to yield a constant number of collisions per unit wall area per unit time.
Figure 7.4.2 : The molecular speed distribution for oxygen gas at 300 K is shown here. Very few molecules move at either very low or very high speeds.
The number of molecules with intermediate speeds increases rapidly up to a maximum, which is the most probable speed, then drops off rapidly. Note that
the most probable speed, νp, is a little less than 400 m/s, while the root mean square speed, urms, is closer to 500 m/s.
The kinetic energy (KE) of a particle of mass (m) and speed (u) is given by:
1
2
KE = mu (7.4.1)
2
Expressing mass in kilograms and speed in meters per second will yield energy values in units of joules (J = kg m2 s–2). To deal with a large number of gas
molecules, we use averages for both speed and kinetic energy. In the KMT, the root mean square velocity of a particle, urms, is defined as the square root
of the average of the squares of the velocities with n = the number of particles:
−−−−−−−−−−−−−−−−−−−
−− 2 2 2 2
¯
¯¯¯
¯ u +u +u +u +…
√ 2 1 2 3 4
urms = u =√ (7.4.2)
n
The KEavg of a collection of gas molecules is also directly proportional to the temperature of the gas and may be described by the equation:
3
KEavg = RT (7.4.4)
2
where R is the gas constant and T is the kelvin temperature. When used in this equation, the appropriate form of the gas constant is 8.314 J mol-1K-1
(8.314 kg m2s–2mol-1K–1). These two separate equations for KEavg may be combined and rearranged to yield a relation between molecular speed and
temperature:
1 3
2
m urms = RT (7.4.5)
2 2
−−−−−
3RT
urms = √ (7.4.6)
m
7.4.2 https://chem.libretexts.org/@go/page/217294
28.0 g
1 kg
× = 0.028 kg/mol (7.4.8)
1 mol 1000 g
Replace the variables and constants in the root-mean-square velocity formula (Equation 7.4.6), replacing Joules with the equivalent kg m2s–2:
−−−−−
3RT
urms = √ (7.4.9)
m
−−−−−−−−−−−−−−−−−−−−−
3(8.314 J/mol K)(303 K)
urms = √ (7.4.10)
(0.028 kg/mol)
Exercise 7.4.1
Calculate the root-mean-square velocity for an oxygen molecule at –23 °C.
Answer
441 m/s
If the temperature of a gas increases, its KEavg increases, more molecules have higher speeds and fewer molecules have lower speeds, and the distribution
shifts toward higher speeds overall, that is, to the right. If temperature decreases, KEavg decreases, more molecules have lower speeds and fewer molecules
have higher speeds, and the distribution shifts toward lower speeds overall, that is, to the left. This behavior is illustrated for nitrogen gas in Figure 7.4.3.
Figure 7.4.3 : The molecular speed distribution for nitrogen gas (N2) shifts to the right and flattens as the temperature increases; it shifts to the left and
heightens as the temperature decreases.
At a given temperature, all gases have the same KEavg for their molecules. Gases composed of lighter molecules have more high-speed particles and a
higher urms, with a speed distribution that peaks at relatively higher velocities. Gases consisting of heavier molecules have more low-speed particles, a
lower urms, and a speed distribution that peaks at relatively lower velocities. This trend is demonstrated by the data for a series of noble gases shown in
Figure 7.4.4.
Figure 7.4.4 : Molecular velocity is directly related to molecular mass. At a given temperature, lighter molecules move faster on average than heavier
molecules.
7.4.3 https://chem.libretexts.org/@go/page/217294
Gas Properties
Ideal
Using this relation, and the equation relating molecular speed to mass, Graham’s law may be easily derived as shown here:
−−−−−
3RT
urms = √ (7.4.14)
m
3RT 3RT
m = = (7.4.15)
2 2
urms ¯¯
ū
−−−−−
3RT
√
−−−
−
ef f usion rate A urms A
mA mB
= = −−−−− =√ (7.4.16)
ef f usion rate B urms B 3RT mA
√
mB
The ratio of the rates of effusion is thus derived to be inversely proportional to the ratio of the square roots of their masses. This is the same relation
observed experimentally and expressed as Graham’s law.
Summary
7.4.4 https://chem.libretexts.org/@go/page/217294
Kinetic Theory and Phase Changes: Crash Cours…
Cours…
Video 7.4.1 : A summary of Kinetic Molecular Theory under the Ideal Gas Law.
The kinetic molecular theory is a simple but very effective model that effectively explains ideal gas behavior. The theory assumes that gases consist of
widely separated molecules of negligible volume that are in constant motion, colliding elastically with one another and the walls of their container with
average velocities determined by their absolute temperatures. The individual molecules of a gas exhibit a range of velocities, the distribution of these
velocities being dependent on the temperature of the gas and the mass of its molecules.
Key Equations
−−−−−−−−−−−−−−−−−−−
−− 2 2 2 2
¯
¯¯¯
¯
u +u +u +u +…
1 2 3 4
√ 2
urms = u =√
n
3
KEavg = RT
2
−−−−−
3RT
urms =√
m
Glossary
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/[email protected]).
Adelaide Clark, Oregon Institute of Techonology
Crash Course Physics: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
7.4: Kinetic Molecular Theory is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
7.4.5 https://chem.libretexts.org/@go/page/217294
7.4.1: Practice Problems- Kinetic Molecular Theory
PROBLEM 7.4.1.1
Using the postulates of the kinetic molecular theory, explain why a gas uniformly fills a container of any shape.
Answer
The first three parts of Kinetic Molecular Theory explain this:
Gases are composed of molecules that are in continuous motion, travelling in straight lines and changing direction only
when they collide with other molecules or with the walls of a container.
The molecules composing the gas are negligibly small compared to the distances between them.
The pressure exerted by a gas in a container results from collisions between the gas molecules and the container walls.
PROBLEM 7.4.1.2
Can the speed of a given molecule in a gas double at constant temperature? Explain your answer.
Answer
Yes. At any given instant, there are a range of values of molecular speeds in a sample of gas. Any single molecule can speed
up or slow down as it collides with other molecules. The average velocity of all the molecules is constant at constant
temperature.
PROBLEM 7.4.1.3
Describe what happens to the average kinetic energy of ideal gas molecules when the conditions are changed as follows:
a. The pressure of the gas is increased by reducing the volume at constant temperature.
b. The pressure of the gas is increased by increasing the temperature at constant volume.
c. The average velocity of the molecules is increased by a factor of 2.
Answer a
Kinetic energy will not change (temperature is unchanged)
Answer b
Kinetic energy will increase (temperature is increased)
Answer c
The kinetic energy will increase by a factor of four (velocity is squared in the equation)
PROBLEM 7.4.1.4
The distribution of molecular velocities in a sample of helium is shown in the photo below (Figure 7.4.4). If the sample is
cooled, will the distribution of velocities look more like that of H2 or of H2O? Explain your answer.
7.4.1.1 https://chem.libretexts.org/@go/page/217295
Answer
H2O. Cooling slows the velocities of the He atoms, causing them to behave as though they were heavier.
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
7.4.1: Practice Problems- Kinetic Molecular Theory is shared under a not declared license and was authored, remixed, and/or curated by
LibreTexts.
7.4.1.2 https://chem.libretexts.org/@go/page/217295
CHAPTER OVERVIEW
8: Thermochemistry
Unit 8 Objectives
Define energy, distinguish types of energy, and describe the nature of energy changes that accompany chemical and physical
changes
Distinguish the related properties of heat, thermal energy, and temperature
Define and distinguish specific heat and heat capacity, and describe the physical implications of both
Perform calculations involving heat, specific heat, and temperature change
State the first law of thermodynamics
Define enthalpy and explain its classification as a state function
Write and balance thermochemical equations
Calculate enthalpy changes for various chemical reactions
Explain Hess’s law and use it to compute reaction enthalpies
Explain the technique of calorimetry
Calculate and interpret heat and related properties using typical calorimetry data
8: Thermochemistry is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
1
8.1: The Basics of Energy
Skills to Develop
Define energy, distinguish types of energy, and describe the nature of energy changes that accompany chemical and physical
changes
Distinguish the related properties of heat, thermal energy, and temperature
Chemical changes and their accompanying changes in energy are important parts of our everyday world (Figure 8.1.1). The
macronutrients in food (proteins, fats, and carbohydrates) undergo metabolic reactions that provide the energy to keep our bodies
functioning. We burn a variety of fuels (gasoline, natural gas, coal) to produce energy for transportation, heating, and the
generation of electricity. Industrial chemical reactions use enormous amounts of energy to produce raw materials (such as iron and
aluminum). Energy is then used to manufacture those raw materials into useful products, such as cars, skyscrapers, and bridges.
Figure 8.1.1 : The energy involved in chemical changes is important to our daily lives: (a) A cheeseburger for lunch provides the
energy you need to get through the rest of the day; (b) the combustion of gasoline provides the energy that moves your car (and
you) between home, work, and school; and (c) coke, a processed form of coal, provides the energy needed to convert iron ore into
iron, which is essential for making many of the products we use daily. (credit a: modification of work by “Pink Sherbet
Photography”/Flickr; credit b: modification of work by Jeffery Turner).
Over 90% of the energy we use comes originally from the sun. Every day, the sun provides the earth with almost 10,000 times the
amount of energy necessary to meet all of the world’s energy needs for that day. Our challenge is to find ways to convert and store
incoming solar energy so that it can be used in reactions or chemical processes that are both convenient and nonpolluting. Plants
and many bacteria capture solar energy through photosynthesis. We release the energy stored in plants when we burn wood or plant
products such as ethanol. We also use this energy to fuel our bodies by eating food that comes directly from plants or from animals
that got their energy by eating plants. Burning coal and petroleum also releases stored solar energy: These fuels are fossilized plant
and animal matter.
This chapter will introduce the basic ideas of an important area of science concerned with the amount of heat absorbed or released
during chemical and physical changes—an area called thermochemistry. The concepts introduced in this chapter are widely used in
almost all scientific and technical fields. Food scientists use them to determine the energy content of foods. Biologists study the
energetics of living organisms, such as the metabolic combustion of sugar into carbon dioxide and water. The oil, gas, and
transportation industries, renewable energy providers, and many others endeavor to find better methods to produce energy for our
commercial and personal needs. Engineers strive to improve energy efficiency, find better ways to heat and cool our homes,
refrigerate our food and drinks, and meet the energy and cooling needs of computers and electronics, among other applications.
Understanding thermochemical principles is essential for chemists, physicists, biologists, geologists, every type of engineer, and
just about anyone who studies or does any kind of science.
Energy
Energy can be defined as the capacity to supply heat or do work. One type of work (w) is the process of causing matter to move
against an opposing force. For example, we do work when we inflate a bicycle tire—we move matter (the air in the pump) against
the opposing force of the air surrounding the tire.
Like matter, energy comes in different types. One scheme classifies energy into two types: potential energy, the energy an object
has because of its relative position, composition, or condition, and kinetic energy, the energy that an object possesses because of its
motion. Water at the top of a waterfall or dam has potential energy because of its position; when it flows downward through
8.1.1 https://chem.libretexts.org/@go/page/217297
generators, it has kinetic energy that can be used to do work and produce electricity in a hydroelectric plant (Figure 8.1.2). A
battery has potential energy because the chemicals within it can produce electricity that can do work.
Figure 8.1.2 : (a) Water that is higher in elevation, for example, at the top of Victoria Falls, has a higher potential energy than
water at a lower elevation. As the water falls, some of its potential energy is converted into kinetic energy. (b) If the water flows
through generators at the bottom of a dam, such as the Hoover Dam shown here, its kinetic energy is converted into electrical
energy. (credit a: modification of work by Steve Jurvetson; credit b: modification of work by “curimedia”/Wikimedia commons).
Energy can be converted from one form into another, but all of the energy present before a change occurs always exists in some
form after the change is completed. This observation is expressed in the law of conservation of energy: during a chemical or
physical change, energy can be neither created nor destroyed, although it can be changed in form. (This is also one version of the
first law of thermodynamics, as you will learn later.)
When one substance is converted into another, there is always an associated conversion of one form of energy into another. Heat is
usually released or absorbed, but sometimes the conversion involves light, electrical energy, or some other form of energy. For
example, chemical energy (a type of potential energy) is stored in the molecules that compose gasoline. When gasoline is
combusted within the cylinders of a car’s engine, the rapidly expanding gaseous products of this chemical reaction generate
mechanical energy (a type of kinetic energy) when they move the cylinders’ pistons.
According to the law of conservation of matter (seen in an earlier chapter), there is no detectable change in the total amount of
matter during a chemical change. When chemical reactions occur, the energy changes are relatively modest and the mass changes
are too small to measure, so the laws of conservation of matter and energy hold well. However, in nuclear reactions, the energy
changes are much larger (by factors of a million or so), the mass changes are measurable, and matter-energy conversions are
significant. This will be examined in more detail in a later chapter on nuclear chemistry. To encompass both chemical and nuclear
changes, we combine these laws into one statement: The total quantity of matter and energy in the universe is fixed.
8.1.2 https://chem.libretexts.org/@go/page/217297
'
Figure 8.1.3 : (a) The molecules in a sample of hot water move more rapidly than (b) those in a sample of cold water.
Phase Change
States
Most substances expand as their temperature increases and contract as their temperature decreases. This property can be used to
measure temperature changes, as shown in Figure 8.1.4. The operation of many thermometers depends on the expansion and
contraction of substances in response to temperature changes.
8.1.3 https://chem.libretexts.org/@go/page/217297
Figure 8.1.4 : (a) In an alcohol or mercury thermometer, the liquid (dyed red for visibility) expands when heated and contracts
when cooled, much more so than the glass tube that contains the liquid. (b) In a bimetallic thermometer, two different metals (such
as brass and steel) form a two-layered strip. When heated or cooled, one of the metals (brass) expands or contracts more than the
other metal (steel), causing the strip to coil or uncoil. Both types of thermometers have a calibrated scale that indicates the
temperature. (credit a: modification of work by “dwstucke”/Flickr). (c) The demonstration allows one to view the effects of heating
and cooling a coiled bimetallic strip.A bimetallic coil from a thermometer reacts to the heat from a lighter, by uncoiling and then
coiling back up when the lighter is removed. Animation used with permission from Hustvedt (via Wikipedia)
Heat (q) is the transfer of thermal energy between two bodies at different temperatures. Heat flow (a redundant term, but one
commonly used) increases the thermal energy of one body and decreases the thermal energy of the other. Suppose we initially have
a high temperature (and high thermal energy) substance (H) and a low temperature (and low thermal energy) substance (L). The
atoms and molecules in H have a higher average KE than those in L. If we place substance H in contact with substance L, the
thermal energy will flow spontaneously from substance H to substance L. The temperature of substance H will decrease, as will the
average KE of its molecules; the temperature of substance L will increase, along with the average KE of its molecules. Heat flow
will continue until the two substances are at the same temperature (Figure 8.1.5).
Figure 8.1.5 : (a) Substances H and L are initially at different temperatures, and their atoms have different average kinetic
energies. (b) When they are put into contact with each other, collisions between the molecules result in the transfer of kinetic
(thermal) energy from the hotter to the cooler matter. (c) The two objects reach “thermal equilibrium” when both substances are at
the same temperature, and their molecules have the same average kinetic energy.
Matter undergoing chemical reactions and physical changes can release or absorb heat. A change that releases heat is called an
exothermic process. For example, the combustion reaction that occurs when using an oxyacetylene torch is an exothermic process
—this process also releases energy in the form of light as evidenced by the torch’s flame (Figure 8.1.6a). A reaction or change that
absorbs heat is an endothermic process. A cold pack used to treat muscle strains provides an example of an endothermic process.
When the substances in the cold pack (water and a salt like ammonium nitrate) are brought together, the resulting process absorbs
heat, leading to the sensation of cold.
8.1.4 https://chem.libretexts.org/@go/page/217297
Figure 8.1.6 : (a) An oxyacetylene torch produces heat by the combustion of acetylene in oxygen. The energy released by this
exothermic reaction heats and then melts the metal being cut. The sparks are tiny bits of the molten metal flying away. (b) A cold
pack uses an endothermic process to create the sensation of cold. (credit a: modification of work by “Skatebiker”/Wikimedia
commons).
Units of Energy
Energy is measured in terms of its ability to perform work or to transfer heat. Mechanical work is done when a force f displaces an
object by a distance d:
w = f ×d (8.1.1)
The basic unit of energy is the joule. One joule is the amount of work done when a force of 1 newton acts over a distance of 1 m;
thus 1 J = 1 N-m. The newton is the amount of force required to accelerate a 1-kg mass by 1 m/sec2, so the basic dimensions of the
joule are kg m2 s–2. Historically, energy was measured in units of calories (cal). A calorie is the amount of energy required to raise
one gram of water by 1 degree C (1 kelvin). However, this quantity depends on the atmospheric pressure and the starting
temperature of the water. The ease of measurement of energy changes in calories has meant that the calorie is still frequently used.
The Calorie (with a capital C), or large calorie, commonly used in quantifying food energy content, is a kilocalorie. Another
common unit of measurement of energy is the BTU (British thermal unit) which is defined in terms of the heating effect on water.
Because of the many forms that energy can take, there are a correspondingly large number of units in which it can be expressed, a
few of which are summarized below.
The erg is the c.g.s. unit of energy and a very small one; the work done 1 J = 107 ergs
when a 1-dyne force acts over a distance of 1 cm. 1 erg = 1 d-cm = 1 g cm2 s–2
The watt is a unit of power, which measures the rate of energy flow in J
sec–1. Thus the watt-hour is a unit of energy. An average human 1 J = 2.78 × 10–4watt-hr
consumes energy at a rate of about 100 watts; the brain alone runs at 1 w-h = 3.6 kJ
about 5 watts.
If the object is to obliterate cities or countries with nuclear weapons, 1 ton of TNT = 4.184 GJ
the energy unit of choice is the ton of TNT equivalent. (by definition)
1 bboe = 6.1 GJ
In terms of fossil fuels, we have barrel-of-oil equivalent, cubic-meter-
1 cmge = 37-39 mJ
of-natural gas equivalent, and ton-of-coal equivalent.
1 toce = 29 GJ
Summary
8.1.5 https://chem.libretexts.org/@go/page/217297
Energy & Chemistry: Crash Course Che…
Che…
Key Equations
q = c × m × ΔT = c × m × (Tfinal − Tinitial)
Glossary
calorie (cal)
unit of heat or other energy; the amount of energy required to raise 1 gram of water by 1 degree Celsius; 1 cal is defined as
4.184 J
endothermic process
chemical reaction or physical change that absorbs heat
energy
capacity to supply heat or do work
exothermic process
chemical reaction or physical change that releases heat
joule (J)
SI unit of energy; 1 joule is the kinetic energy of an object with a mass of 2 kilograms moving with a velocity of 1 meter per
second, 1 J = 1 kg m2/s and 4.184 J = 1 cal
kinetic energy
1
energy of a moving body, in joules, equal to 2
mv (where m = mass and v = velocity)
2
potential energy
energy of a particle or system of particles derived from relative position, composition, or condition
temperature
8.1.6 https://chem.libretexts.org/@go/page/217297
intensive property of matter that is a quantitative measure of “hotness” and “coldness”
thermal energy
kinetic energy associated with the random motion of atoms and molecules
thermochemistry
study of measuring the amount of heat absorbed or released during a chemical reaction or a physical change
work (w)
energy transfer due to changes in external, macroscopic variables such as pressure and volume; or causing matter to move
against an opposing force
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide E. Clark, Oregon Institute of Technology
Stephen Lower, Professor Emeritus (Simon Fraser U.) Chem1 Virtual Textbook
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
8.1: The Basics of Energy is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
8.1.7 https://chem.libretexts.org/@go/page/217297
8.1.1: Practice Problems- The Basics of Energy
PROBLEM 8.1.1.1
A burning match and a bonfire may have the same temperature, yet you would not sit around a burning match on a fall evening
to stay warm. Why not?
Answer
The temperature of 1 gram of burning wood is approximately the same for both a match and a bonfire. This is an intensive
property and depends on the material (wood). However, the overall amount of produced heat depends on the amount of
material; this is an extensive property. The amount of wood in a bonfire is much greater than that in a match; the total
amount of produced heat is also much greater, which is why we can sit around a bonfire to stay warm, but a match would not
provide enough heat to keep us from getting cold.
PROBLEM 8.1.1.2
Explain the difference between heat capacity and specific heat of a substance.
Answer
Heat capacity refers to the heat required to raise the temperature of the mass of the substance 1 degree; specific heat refers to
the heat required to raise the temperature of 1 gram of the substance 1 degree. Thus, heat capacity is an extensive property,
and specific heat is an intensive one.
PROBLEM 8.1.1.3
How much heat, in joules and in calories, must be added to a 75.0–g iron block with a specific heat of 0.449 J/g °C to increase
its temperature from 25 °C to its melting temperature of 1535 °C?
Answer
50,800 J
12,200 cal
Problem 8.1.3
8.1.1.1 https://chem.libretexts.org/@go/page/217298
PROBLEM 8.1.1.4
How much heat, in joules and in calories, is required to heat a 28.4-g (1-oz) ice cube from −23.0 °C to −1.0 °C?
Answer
1310 J
313 cal
PROBLEM 8.1.1.5
How much would the temperature of 275 g of water increase if 36.5 kJ of heat were added?
Answer
31.7° C
Problem 8.1.5
PROBLEM 8.1.1.6
If 14.5 kJ of heat were added to 485 g of liquid water, how much would its temperature increase?
Answer
7.15 °C
PROBLEM 8.1.1.7
A piece of unknown substance weighs 44.7 g and requires 2110 J to increase its temperature from 23.2 °C to 89.6 °C.
a. What is the specific heat of the substance?
b. If it is one of the substances found in Table 8.1.1, what is its likely identity?
Answer a
0.711 J
C =
g °C
8.1.1.2 https://chem.libretexts.org/@go/page/217298
Answer b
Silicon
Problem 8.1.7
PROBLEM 8.1.1.8
A piece of unknown solid substance weighs 437.2 g, and requires 8460 J to increase its temperature from 19.3 °C to 68.9 °C.
a. What is the specific heat of the substance?
b. If it is one of the substances found in Table 8.1.1, what is its likely identity?
Answer a
0.390 J
C =
g °C
Answer b
Copper
PROBLEM 8.1.1.9
An aluminum kettle weighs 1.05 kg.
a. What is the heat capacity of the kettle (Table 8.1.1)?
b. How much heat is required to increase the temperature of this kettle from 23.0 °C to 99.0 °C?
c. How much heat (in kJ) is required to heat this kettle from 23.0 °C to 99.0 °C if it contains 1.25 L of water (density of 0.997
g/mL and a specific heat of 4.184 J/g °C)?
Answer a
0.897 J
C =
g °C
Answer b
71580 J
8.1.1.3 https://chem.libretexts.org/@go/page/217298
Answer c
467.86 kJ
Problem 8.1.9
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
8.1.1: Practice Problems- The Basics of Energy is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
8.1.1.4 https://chem.libretexts.org/@go/page/217298
8.2: Measuring Energy and Heat Capacity
Skills to Develop
Define and distinguish specific heat and heat capacity, and describe the physical implications of both
Perform calculations involving heat, specific heat, and temperature change
Heat capacity is determined by both the type and amount of substance that absorbs or releases heat. It is therefore an extensive
property—its value is proportional to the amount of the substance.
For example, consider the heat capacities of two cast iron frying pans. The heat capacity of the large pan is five times greater than
that of the small pan because, although both are made of the same material, the mass of the large pan is five times greater than the
mass of the small pan. More mass means more atoms are present in the larger pan, so it takes more energy to make all of those
atoms vibrate faster. The heat capacity of the small cast iron frying pan is found by observing that it takes 18,150 J of energy to
raise the temperature of the pan by 50.0 °C
18, 140 J
Csmall pan = = 363 J/°C (8.2.2)
50.0 °C
The larger cast iron frying pan, while made of the same substance, requires 90,700 J of energy to raise its temperature by 50.0 °C.
The larger pan has a (proportionally) larger heat capacity because the larger amount of material requires a (proportionally) larger
amount of energy to yield the same temperature change:
90, 700 J
Clarge pan = = 1814 J/°C (8.2.3)
50.0 °C
The specific heat capacity (c) of a substance, commonly called its “specific heat,” is the quantity of heat required to raise the
temperature of 1 gram of a substance by 1 degree Celsius (or 1 kelvin):
q
c = (8.2.4)
mΔT
Specific heat capacity depends only on the kind of substance absorbing or releasing heat. It is an intensive property—the type, but
not the amount, of the substance is all that matters. For example, the small cast iron frying pan has a mass of 808 g. The specific
heat of iron (the material used to make the pan) is therefore:
18, 140 J
ciron = = 0.449 J/g °C (8.2.5)
(808 g)(50.0 °C)
The large frying pan has a mass of 4040 g. Using the data for this pan, we can also calculate the specific heat of iron:
90, 700 J
ciron = = 0.449 J/g °C (8.2.6)
(4, 040 g)(50.0 °C)
Although the large pan is more massive than the small pan, since both are made of the same material, they both yield the same
value for specific heat (for the material of construction, iron). Note that specific heat is measured in units of energy per temperature
per mass and is an intensive property, being derived from a ratio of two extensive properties (heat and mass). The molar heat
capacity, also an intensive property, is the heat capacity per mole of a particular substance and has units of J/mol °C (Figure 8.2.7).
8.2.1 https://chem.libretexts.org/@go/page/222607
Figure 8.2.1 : Due to its larger mass, a large frying pan has a larger heat capacity than a small frying pan. Because they are
made of the same material, both frying pans have the same specific heat. (credit: Mark Blaser).
Liquid water has a relatively high specific heat (about 4.2 J/g °C); most metals have much lower specific heats (usually less than 1
J/g °C). The specific heat of a substance varies somewhat with temperature. However, this variation is usually small enough that we
will treat specific heat as constant over the range of temperatures that will be considered in this chapter. Specific heats of some
common substances are listed in Table 8.2.1.
Table 8.2.1: Specific Heats of Common Substances at 25 °C and 1 bar
Substance Symbol (state) Specific Heat (J/g °C)
air 1.007
If we know the mass of a substance and its specific heat, we can determine the amount of heat, q, entering or leaving the substance
by measuring the temperature change before and after the heat is gained or lost:
q = c × m × ΔT = c × m × (Tfinal − Tinitial)
In this equation, c is the specific heat of the substance, m is its mass, and ΔT (which is read “delta T”) is the temperature change,
Tfinal − Tinitial. If a substance gains thermal energy, its temperature increases, its final temperature is higher than its initial
temperature, Tfinal − Tinitial has a positive value, and the value of q is positive. If a substance loses thermal energy, its temperature
decreases, the final temperature is lower than the initial temperature, Tfinal − Tinitial has a negative value, and the value of q is
negative.
8.2.2 https://chem.libretexts.org/@go/page/222607
Example 8.2.1 : Measuring Heat
A flask containing 8.0 × 10 2
g of water is heated, and the temperature of the water increases from 21 °C to 85 °C. How much
heat did the water absorb?
Solution
To answer this question, consider these factors:
the specific heat of the substance being heated (in this case, water)
the amount of substance being heated (in this case, 800 g)
the magnitude of the temperature change (in this case, from 21 °C to 85 °C).
The specific heat of water is 4.184 J/g °C, so to heat 1 g of water by 1 °C requires 4.184 J. We note that since 4.184 J is required
to heat 1 g of water by 1 °C, we will need 800 times as much to heat 800 g of water by 1 °C. Finally, we observe that since
4.184 J are required to heat 1 g of water by 1 °C, we will need 64 times as much to heat it by 64 °C (that is, from 21 °C to 85
°C).
This can be summarized using the equation:
q = c × m × ΔT = c × m × (Tfinal − Tinitial)
Because the temperature increased, the water absorbed heat and q is positive.
Exercise 8.2.1
How much heat, in joules, must be added to a 5.00 × 10 2
g iron skillet to increase its temperature from 25 °C to 250 °C? The
specific heat of iron is 0.451 J/g °C.
Answer
4
5.05 × 10 J
Note that the relationship between heat, specific heat, mass, and temperature change can be used to determine any of these
quantities (not just heat) if the other three are known or can be deduced.
Solving:
6, 640 J
c = = 0.900 J/g °C (8.2.9)
(348 g) × (21.2°C)
Comparing this value with the values in Table 8.2.1, this value matches the specific heat of aluminum, which suggests that the
unknown metal may be aluminum.
8.2.3 https://chem.libretexts.org/@go/page/222607
Exercise 8.2.2
A piece of unknown metal weighs 217 g. When the metal piece absorbs 1.43 kJ of heat, its temperature increases from 24.5 °C
to 39.1 °C. Determine the specific heat of this metal, and predict its identity.
Answer
c = 0.45 J/g °C ; the metal is likely to be iron from checking Table 8.2.1.
Figure 8.2.2 : This solar thermal plant uses parabolic trough mirrors to concentrate sunlight. (credit a: modification of work
by Bureau of Land Management)
The 377-megawatt Ivanpah Solar Generating System, located in the Mojave Desert in California, is the largest solar thermal
power plant in the world (Figure 8.2.9). Its 170,000 mirrors focus huge amounts of sunlight on three water-filled towers,
producing steam at over 538 °C that drives electricity-producing turbines. It produces enough energy to power 140,000 homes.
Water is used as the working fluid because of its large heat capacity and heat of vaporization.
Figure 8.2.3 : (a) The Ivanpah solar thermal plant uses 170,000 mirrors to concentrate sunlight on water-filled towers. (b) It
covers 4000 acres of public land near the Mojave Desert and the California-Nevada border. (credit a: modification of work by
Craig Dietrich; credit b: modification of work by “USFWS Pacific Southwest Region”/Flickr)
8.2.4 https://chem.libretexts.org/@go/page/222607
Summary
Key Equations
q = c × m × ΔT = c × m × (Tfinal − Tinitial)
Glossary
calorie (cal)
unit of heat or other energy; the amount of energy required to raise 1 gram of water by 1 degree Celsius; 1 cal is defined as
4.184 J
endothermic process
chemical reaction or physical change that absorbs heat
energy
capacity to supply heat or do work
exothermic process
chemical reaction or physical change that releases heat
heat (q)
transfer of thermal energy between two bodies
joule (J)
SI unit of energy; 1 joule is the kinetic energy of an object with a mass of 2 kilograms moving with a velocity of 1 meter per
second, 1 J = 1 kg m2/s and 4.184 J = 1 cal
8.2.5 https://chem.libretexts.org/@go/page/222607
kinetic energy
1
energy of a moving body, in joules, equal to mv
2
(where m = mass and v = velocity)
2
potential energy
energy of a particle or system of particles derived from relative position, composition, or condition
temperature
intensive property of matter that is a quantitative measure of “hotness” and “coldness”
thermal energy
kinetic energy associated with the random motion of atoms and molecules
thermochemistry
study of measuring the amount of heat absorbed or released during a chemical reaction or a physical change
work (w)
energy transfer due to changes in external, macroscopic variables such as pressure and volume; or causing matter to move
against an opposing force
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide E. Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
8.2: Measuring Energy and Heat Capacity is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
8.2.6 https://chem.libretexts.org/@go/page/222607
8.3: Enthalpy
Skills to Develop
State the first law of thermodynamics
Define enthalpy and explain its classification as a state function
Thermochemistry is a branch of chemical thermodynamics, the science that deals with the relationships between heat, work, and
other forms of energy in the context of chemical and physical processes. As we concentrate on thermochemistry in this chapter, we
need to consider some widely used concepts of thermodynamics.
Substances act as reservoirs of energy, meaning that energy can be added to them or removed from them. Energy is stored in a
substance when the kinetic energy of its atoms or molecules is raised. The greater kinetic energy may be in the form of increased
translations (travel or straight-line motions), vibrations, or rotations of the atoms or molecules. When thermal energy is lost, the
intensities of these motions decrease and the kinetic energy falls. The total of all possible kinds of energy present in a substance is
called the internal energy (U), sometimes symbolized as E.
As a system undergoes a change, its internal energy can change, and energy can be transferred from the system to the surroundings,
or from the surroundings to the system. Energy is transferred into a system when it absorbs heat (q) from the surroundings or when
the surroundings do work (w) on the system. For example, energy is transferred into room-temperature metal wire if it is immersed
in hot water (the wire absorbs heat from the water), or if you rapidly bend the wire back and forth (the wire becomes warmer
because of the work done on it). Both processes increase the internal energy of the wire, which is reflected in an increase in the
wire’s temperature. Conversely, energy is transferred out of a system when heat is lost from the system, or when the system does
work on the surroundings.
The relationship between internal energy, heat, and work can be represented by the equation:
ΔU = q + w (8.3.1)
as shown in Figure 8.3.1. This is one version of the first law of thermodynamics, and it shows that the internal energy of a system
changes through heat flow into or out of the system (positive q is heat flow in; negative q is heat flow out) or work done on or by
the system. The work, w, is positive if it is done on the system and negative if it is done by the system.
Figure 8.3.1 : The internal energy, U, of a system can be changed by heat flow and work. If heat flows into the system, qin, or work
is done on the system, won, its internal energy increases, ΔU > 0. If heat flows out of the system, qout, or work is done by the system,
wby, its internal energy decreases, ΔU < 0.
A type of work called expansion work (or pressure-volume work) occurs when a system pushes back the surroundings against a
restraining pressure, or when the surroundings compress the system. An example of this occurs during the operation of an internal
combustion engine. The reaction of gasoline and oxygen is exothermic. Some of this energy is given off as heat, and some does
work pushing the piston in the cylinder. The substances involved in the reaction are the system, and the engine and the rest of the
universe are the surroundings. The system loses energy by both heating and doing work on the surroundings, and its internal energy
decreases. (The engine is able to keep the car moving because this process is repeated many times per second while the engine is
running.) We will consider how to determine the amount of work involved in a chemical or physical change in the chapter on
thermodynamics.
8.3.1 https://chem.libretexts.org/@go/page/217301
As discussed, the relationship between internal energy, heat, and work can be represented as ΔU = q + w. Internal energy is a type
of quantity known as a state function (or state variable), whereas heat and work are not state functions. The value of a state function
depends only on the state that a system is in, and not on how that state is reached. If a quantity is not a state function, then its value
does depend on how the state is reached. An example of a state function is altitude or elevation. If you stand on the summit of Mt.
Kilimanjaro, you are at an altitude of 5895 m, and it does not matter whether you hiked there or parachuted there. The distance you
traveled to the top of Kilimanjaro, however, is not a state function. You could climb to the summit by a direct route or by a more
roundabout, circuitous path (Figure 8.3.2). The distances traveled would differ (distance is not a state function) but the elevation
reached would be the same (altitude is a state function).
Figure 8.3.2 : Paths X and Y represent two different routes to the summit of Mt. Kilimanjaro. Both have the same change in
elevation (altitude or elevation on a mountain is a state function; it does not depend on path), but they have very different distances
traveled (distance walked is not a state function; it depends on the path). (credit: modification of work by Paul Shaffner)
Chemists ordinarily use a property known as enthalpy (H ) to describe the thermodynamics of chemical and physical processes.
Enthalpy is defined as the sum of a system’s internal energy (U ) and the mathematical product of its pressure (P ) and volume (V ):
H = U +PV (8.3.2)
Since it is derived from three state functions (U , P , and V ), enthalpy is also a state function. Enthalpy values for specific
substances cannot be measured directly; only enthalpy changes for chemical or physical processes can be determined. For
processes that take place at constant pressure (a common condition for many chemical and physical changes), the enthalpy change (
ΔH ) is:
ΔH = ΔU + P ΔV (8.3.3)
The mathematical product P ΔV represents work (w), namely, expansion or pressure-volume work as noted. By their definitions,
the arithmetic signs of ΔV and w will always be opposite:
P ΔV = −w (8.3.4)
Substituting Equation 8.3.4 and the definition of internal energy (Equation 8.3.1) into Equation 8.3.3 yields:
ΔH = ΔU + P ΔV (8.3.5)
= qp + w − w (8.3.6)
= qp (8.3.7)
And so, if a chemical or physical process is carried out at constant pressure with the only work done caused by expansion or
contraction, then the heat flow (q ) and enthalpy change (ΔH ) for the process are equal.
p
The heat given off when you operate a Bunsen burner is equal to the enthalpy change of the methane combustion reaction that takes
place, since it occurs at the essentially constant pressure of the atmosphere. On the other hand, the heat produced by a reaction
measured in a bomb calorimeter is not equal to ΔH because the closed, constant-volume metal container prevents expansion work
from occurring. Chemists usually perform experiments under normal atmospheric conditions, at constant external pressure with
q = ΔH , which makes enthalpy the most convenient choice for determining heat.
8.3.2 https://chem.libretexts.org/@go/page/217301
1. Chemists use a thermochemical equation to represent the changes in both matter and energy. In a thermochemical equation, the
enthalpy change of a reaction is shown as a ΔH value following the equation for the reaction. This ΔH value indicates the
amount of heat associated with the reaction involving the number of moles of reactants and products as shown in the chemical
equation. For example, consider this equation:
1
H (g) + O (g) ⟶ H O(l) ΔH = −286 kJ (8.3.8)
2 2 2 2
This equation indicates that when 1 mole of hydrogen gas and 12 mole of oxygen gas at some temperature and pressure change
to 1 mole of liquid water at the same temperature and pressure, 286 kJ of heat are released to the surroundings. If the
coefficients of the chemical equation are multiplied by some factor, the enthalpy change must be multiplied by that same factor
(ΔH is an extensive property).
(two-fold increase in amounts)
1 1 1 1
H (g) + O (g) ⟶ H O(l) ΔH = × (−286 kJ) = −143 kJ
2 2 2
2 4 2 2
2. The enthalpy change of a reaction depends on the physical state of the reactants and products of the reaction (whether we have
gases, liquids, solids, or aqueous solutions), so these must be shown. For example, when 1 mole of hydrogen gas and 12 mole
of oxygen gas change to 1 mole of liquid water at the same temperature and pressure, 286 kJ of heat are released. If gaseous
water forms, only 242 kJ of heat are released.
1
H (g) + O (g) ⟶ H O(g) ΔH = −242 kJ (8.3.9)
2 2 2 2
3. A negative value of an enthalpy change, ΔH, indicates an exothermic reaction; a positive value of ΔH indicates an endothermic
reaction. If the direction of a chemical equation is reversed, the arithmetic sign of its ΔH is changed (a process that is
endothermic in one direction is exothermic in the opposite direction).
Solution
For the reaction of 0.0500 mol acid (HCl), q = −2.9 kJ. This ratio
−2.9 kJ
can be used as a conversion factor to find the heat produced when 1 mole of HCl reacts:
−2.9 kJ
ΔH = 1 mol HCl × = −58 kJ
0.0500 mol HCl
The enthalpy change when 1 mole of HCl reacts is −58 kJ. Since that is the number of moles in the chemical equation, we write
the thermochemical equation as:
Exercise 8.3.1
When 1.34 g Zn(s) reacts with 60.0 mL of 0.750 M HCl(aq), 3.14 kJ of heat are produced. Determine the enthalpy change per
mole of zinc reacting for the reaction:
Answer
8.3.3 https://chem.libretexts.org/@go/page/217301
ΔH = −153 kJ
Be sure to take both stoichiometry and limiting reactants into account when determining the ΔH for a chemical reaction.
Solution
1 mol
We have 2.67 g × = 0.00780 mol C12 H22 O11 available, and
342.3 g
1 mol
7.19 g × = 0.0587 mol KCl O3 available.
122.5 g
Since
1 mol C H O
12 22 11
0.0587 mol KCl O3 × = 0.00734 mol C H O
12 22 11
8 mol KClO3
is needed, C12H22O11 is the excess reactant and KClO3 is the limiting reactant.
−43.7 kJ
The reaction uses 8 mol KClO3, and the conversion factor is , so we have
0.0587 mol KClO3
−43.7 kJ
ΔH = 8 mol × = −5960 kJ . The enthalpy change for this reaction is −5960 kJ, and the thermochemical
0.0587 mol KClO3
equation is:
Exercise 8.3.2
When 1.42 g of iron reacts with 1.80 g of chlorine, 3.22 g of FeCl2(s) and 8.60 kJ of heat is produced. What is the enthalpy
change for the reaction when 1 mole of FeCl (s) is produced? 2
Answer
ΔH = −338 kJ
Enthalpy changes are typically tabulated for reactions in which both the reactants and products are at the same conditions. A
standard state is a commonly accepted set of conditions used as a reference point for the determination of properties under other
different conditions. For chemists, the IUPAC standard state refers to materials under a pressure of 1 bar and solutions at 1 M, and
does not specify a temperature (it used too). Many thermochemical tables list values with a standard state of 1 atm. Because the
ΔH of a reaction changes very little with such small changes in pressure (1 bar = 0.987 atm), ΔH values (except for the most
precisely measured values) are essentially the same under both sets of standard conditions. We will include a superscripted “o” in
the enthalpy change symbol to designate standard state. Since the usual (but not technically standard) temperature is 298.15 K, we
will use a subscripted “298” to designate this temperature. Thus, the symbol (ΔH ) is used to indicate an enthalpy change for a
∘
298
process occurring under these conditions. (The symbol ΔH is used to indicate an enthalpy change for a reaction occurring under
nonstandard conditions.)
The enthalpy changes for many types of chemical and physical processes are available in the reference literature, including those
for combustion reactions, phase transitions, and formation reactions. As we discuss these quantities, it is important to pay attention
to the extensive nature of enthalpy and enthalpy changes. Since the enthalpy change for a given reaction is proportional to the
amounts of substances involved, it may be reported on that basis (i.e., as the ΔH for specific amounts of reactants). However, we
often find it more useful to divide one extensive property (ΔH) by another (amount of substance), and report a per-amount intensive
value of ΔH, often “normalized” to a per-mole basis. (Note that this is similar to determining the intensive property specific heat
from the extensive property heat capacity, as seen previously.)
8.3.4 https://chem.libretexts.org/@go/page/217301
Enthalpy of Combustion
Standard enthalpy of combustion (ΔH ) is the enthalpy change when 1 mole of a substance burns (combines vigorously with
∘
C
oxygen) under standard state conditions; it is sometimes called “heat of combustion.” For example, the enthalpy of combustion of
ethanol, −1366.8 kJ/mol, is the amount of heat produced when one mole of ethanol undergoes complete combustion at 25 °C and 1
atmosphere pressure, yielding products also at 25 °C and 1 atm.
∘
C H OH(l) + 3 O (g) ⟶ 2 CO + 3 H O(l) ΔH = −1366.8 kJ (8.3.10)
2 5 2 2 2 298
Enthalpies of combustion for many substances have been measured; a few of these are listed in Table 8.3.1. Many readily available
substances with large enthalpies of combustion are used as fuels, including hydrogen, carbon (as coal or charcoal), and
hydrocarbons (compounds containing only hydrogen and carbon), such as methane, propane, and the major components of
gasoline.
Table 8.3.1: Standard Molar Enthalpies of Combustion
Enthalpy of Combustion
Substance Combustion Reaction ∘
kJ
ΔHc ( at 25°C)
mol
hydrogen H (g) +
2
1
2
O (g) ⟶ H O(l)
2 2
−285.8
magnesium Mg(s) +
1
2
O (g) ⟶ MgO(s)
2
−601.6
2
O (g) ⟶ CO (g)
2 2
−283.0
5
acetylene C H (g) +
2 2
O (g) ⟶ 2 CO (g) + H O(l)
2 2 2
−1301.1
2
3
methanol CH OH(l) +
3
O (g) ⟶ CO (g) + 2 H O(l)
2 2 2
−726.1
2
25
isooctane C H
8 18
(l) + O (g) ⟶ 8 CO (g) + 9 H O(l)
2 2 2
−5461
2
Figure 8.3.3 : The combustion of gasoline is very exothermic. (credit: modification of work by “AlexEagle”/Flickr)
Solution
8.3.5 https://chem.libretexts.org/@go/page/217301
Starting with a known amount (1.00 L of isooctane), we can perform conversions between units until we arrive at the desired
amount of heat or energy. The enthalpy of combustion of isooctane provides one of the necessary conversions. Table 8.3.1 gives
this value as −5460 kJ per 1 mole of isooctane (C8H18).
Using these data,
The combustion of 1.00 L of isooctane produces 33,100 kJ of heat. (This amount of energy is enough to melt 99.2 kg, or about
218 lbs, of ice.)
Note: If you do this calculation one step at a time, you would find:
3
1.00 L C H ⟶ 1.00 × 10 mL C H
8 18 8 18
3
1.00 × 10 mL C H ⟶ 692 g C H
8 18 8 18
4
692 g C H ⟶ −3.31 × 10 kJ
8 18
Exercise 8.3.3
How much heat is produced by the combustion of 125 g of acetylene?
Answer
6.25 × 103 kJ
Figure 8.3.4 : (a) Tiny algal organisms can be (b) grown in large quantities and eventually (c) turned into a useful fuel such as
biodiesel. (credit a: modification of work by Micah Sittig; credit b: modification of work by Robert Kerton; credit c:
modification of work by John F. Williams)
According to the US Department of Energy, only 39,000 square kilometers (about 0.4% of the land mass of the US or less than
1
of the area used to grow corn) can produce enough algal fuel to replace all the petroleum-based fuel used in the US. The cost
7
of algal fuels is becoming more competitive—for instance, the US Air Force is producing jet fuel from algae at a total cost of
under $5 per gallon. The process used to produce algal fuel is as follows: grow the algae (which use sunlight as their energy
source and CO2 as a raw material); harvest the algae; extract the fuel compounds (or precursor compounds); process as
necessary (e.g., perform a transesterification reaction to make biodiesel); purify; and distribute (Figure 8.3.5).
8.3.6 https://chem.libretexts.org/@go/page/217301
Figure 8.3.5 : Algae convert sunlight and carbon dioxide into oil that is harvested, extracted, purified, and transformed into a
variety of renewable fuels.
Summary
Video 8.3.1 : A summary of Enthalpy and Hess' Law from Crash Course Chemistry.
If a chemical change is carried out at constant pressure and the only work done is caused by expansion or contraction, q for the
change is called the enthalpy change with the symbol ΔH, or ΔH ∘
298
for reactions occurring under standard state conditions. The
value of ΔH for a reaction in one direction is equal in magnitude, but opposite in sign, to ΔH for the reaction in the opposite
direction, and ΔH is directly proportional to the quantity of reactants and products. Examples of enthalpy changes include enthalpy
of combustion, enthalpy of fusion, enthalpy of vaporization, and standard enthalpy of formation.
Key Equations
ΔU = q + w
Footnotes
1. For more on algal fuel, see http://www.theguardian.com/environme...n-fuel-problem.
Glossary
chemical thermodynamics
area of science that deals with the relationships between heat, work, and all forms of energy associated with chemical and
physical processes
enthalpy (H)
sum of a system’s internal energy and the mathematical product of its pressure and volume
8.3.7 https://chem.libretexts.org/@go/page/217301
expansion work (pressure-volume work)
work done as a system expands or contracts against external pressure
hydrocarbon
compound composed only of hydrogen and carbon; the major component of fossil fuels
state function
property depending only on the state of a system, and not the path taken to reach that state
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide E. Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
8.3: Enthalpy is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
8.3.8 https://chem.libretexts.org/@go/page/217301
8.4: Standard Enthalpy and Hess’ Law
Skills to Develop
Calculate enthalpy changes for various chemical reactions
Explain Hess’s law and use it to compute reaction enthalpies
from free elements in their most stable states under standard state conditions. These values are especially useful for computing or
predicting enthalpy changes for chemical reactions that are impractical or dangerous to carry out, or for processes for which it is
difficult to make measurements. If we have values for the appropriate standard enthalpies of formation, we can determine the
enthalpy change for any reaction, which we will practice in the next section on Hess’s law.
The standard enthalpy of formation of CO2(g) is −393.5 kJ/mol. This is the enthalpy change for the exothermic reaction:
∘ ∘
C(s) + O (g) ⟶ CO (g) ΔH = ΔH = −393.5 kJ (8.4.1)
2 2 f 298
starting with the reactants at a pressure of 1 atm and 25 °C (with the carbon present as graphite, the most stable form of carbon
under these conditions) and ending with one mole of CO2, also at 1 atm and 25 °C. For nitrogen dioxide, NO , ΔH is 33.2 2(g)
∘
f
A reaction equation with mole of N2 and 1 mole of O2 is correct in this case because the standard enthalpy of formation always
1
∘
3 O (g) ⟶ 2 O (g) ΔH = +286 kJ
2 3 298
Solution
ΔH
f
∘
is the enthalpy change for the formation of one mole of a substance in its standard state from the elements in their
standard states. Thus, ΔH for O3(g) is the enthalpy change for the reaction:
∘
f
3
O (g) ⟶ O (g)
2 3
2
286 kJ
For the formation of 2 mol of O3(g), ΔH ∘
298
= +286 kJ . This ratio, ( ), can be used as a conversion factor to find
2 mol O3
the heat produced when 1 mole of O3(g) is formed, which is the enthalpy of formation for O3(g):
286 kJ
∘
ΔH for 1 mole of O (g) = 1 mol O3 × = 143 kJ
3
2 mol O3
Therefore, ΔH f
∘
[ O (g)] = +143 kJ/mol
3
.
8.4.1 https://chem.libretexts.org/@go/page/222609
Exercise 8.4.1
Hydrogen gas, H2, reacts explosively with gaseous chlorine, Cl2, to form hydrogen chloride, HCl(g). What is the enthalpy
change for the reaction of 1 mole of H2(g) with 1 mole of Cl2(g) if both the reactants and products are at standard state
conditions? The standard enthalpy of formation of HCl(g) is −92.3 kJ/mol.
Answer
For the reaction
∘
H (g) + Cl (g) ⟶ 2 HCl(g) ΔH = −184.6 kJ
2 2 298
b. Ca (PO )
3 4 2(s)
Solution
Remembering that ΔH reaction equations are for forming 1 mole of the compound from its constituent elements under
∘
f
2 2 2 5
2 4 2 3 4 2
Exercise 8.4.2
Write the heat of formation reaction equations for:
a. C H OC
2 5 2
H5(l)
b. Na CO2 3(s)
Answer a
4 C(s, graphite) + 5 H (g) +
2
1
2
O (g) ⟶ C H OC H (l)
2 2 5 2 5
;
Answer b
3
2 Na(s) + C(s, graphite) + O (g) ⟶ Na CO (s)
2 2 3
2
Hess’s Law
There are two ways to determine the amount of heat involved in a chemical change: measure it experimentally, or calculate it from
other experimentally determined enthalpy changes. Some reactions are difficult, if not impossible, to investigate and make accurate
measurements for experimentally. And even when a reaction is not hard to perform or measure, it is convenient to be able to
determine the heat involved in a reaction without having to perform an experiment.
This type of calculation usually involves the use of Hess’s law, which states: If a process can be written as the sum of several
stepwise processes, the enthalpy change of the total process equals the sum of the enthalpy changes of the various steps. Hess’s law
is valid because enthalpy is a state function: Enthalpy changes depend only on where a chemical process starts and ends, but not on
the path it takes from start to finish. For example, we can think of the reaction of carbon with oxygen to form carbon dioxide as
occurring either directly or by a two-step process. The direct process is written:
∘
C(s) + O2(g) ⟶ CO2(g) ΔH = −394 kJ (8.4.3)
298
8.4.2 https://chem.libretexts.org/@go/page/222609
1
∘
C(s) + O2(g) ⟶ CO(g) ΔH = −111 kJ (8.4.4)
298
2
The equation describing the overall reaction is the sum of these two chemical changes:
1
Step 1: C(s) + O (g) ⟶ CO(g)
2
2
1
Step 2: CO(g) + O (g) ⟶ CO (g)
2 2
2
––––––––––––––––––––––––––––––––––––
1 1
Sum: C(s) + O (g) + CO(g) + O (g) ⟶ CO(g) + CO (g)
2 2 2
2 2
Because the CO produced in Step 1 is consumed in Step 2, the net change is:
C(s) + O2(g) ⟶ CO2(g) (8.4.6)
According to Hess’s law, the enthalpy change of the reaction will equal the sum of the enthalpy changes of the steps. We can apply
the data from the experimental enthalpies of combustion in Table 8.4.1 to find the enthalpy change of the entire reaction from its
two steps:
1
∘
C(s) + O (g) ⟶ CO(g) ΔH = −111 kJ
2 298
2
1
∘
CO(g) + O (g) ⟶ CO (g) ΔH = −283 kJ
2 2 298
2
¯
¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯
¯ ¯
¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯
¯
∘
C(s) + O (g) ⟶ CO (g) ΔH = −394 kJ
2 2 298
The result is shown in Figure 8.4.6. We see that ΔH of the overall reaction is the same whether it occurs in one step or two. This
finding (overall ΔH for the reaction = sum of ΔH values for reaction “steps” in the overall reaction) is true in general for chemical
and physical processes.
Figure 8.4.2 : The formation of CO2(g) from its elements can be thought of as occurring in two steps, which sum to the overall
reaction, as described by Hess’s law. The horizontal blue lines represent enthalpies. For an exothermic process, the products are at
lower enthalpy than are the reactants.
Before we further practice using Hess’s law, let us recall two important features of ΔH.
1. ΔH is directly proportional to the quantities of reactants or products. For example, the enthalpy change for the reaction forming
1 mole of NO2(g) is +33.2 kJ:
1
N (g) + O (g) ⟶ NO (g) ΔH = +33.2 kJ (8.4.7)
2 2 2
2
8.4.3 https://chem.libretexts.org/@go/page/222609
When 2 moles of NO2 (twice as much) are formed, the ΔH will be twice as large:
N (g) + 2 O (g) ⟶ 2 NO (g) ΔH = +66.4 kJ (8.4.8)
2 2 2
In general, if we multiply or divide an equation by a number, then the enthalpy change should also be multiplied or divided by
the same number.
2. ΔH for a reaction in one direction is equal in magnitude and opposite in sign to ΔH for the reaction in the reverse direction. For
example, given that:
Then, for the “reverse” reaction, the enthalpy change is also “reversed”:
Using Hess’s Law Determine the enthalpy of formation, ΔH , of FeCl3(s) from the enthalpy changes of the following two-step
f
∘
1
FeCl (s) + Cl (g) ⟶ FeCl (s) ΔH ° = −57.7 kJ
2 2 3
2
Solution
We are trying to find the standard enthalpy of formation of FeCl3(s), which is equal to ΔH° for the reaction:
3
∘
Fe(s) + Cl (g) ⟶ FeCl (s) ΔH =?
2 3 f
2
Looking at the reactions, we see that the reaction for which we want to find ΔH° is the sum of the two reactions with known ΔH
values, so we must sum their ΔHs:
Fe(s) + Cl (g) ⟶ FeCl (s) ΔH ° = −341.8 kJ
2 2
1
FeCl (s) + Cl (g) ⟶ FeCl (s) ΔH ° = −57.7 kJ
2 2 3
2
––––––––––––––––––––––––––––––––––––––––––––––––––––
1
Fe(s) + Cl (g) ⟶ FeCl (s) ΔH ° = −399.5 kJ
2 3
2
Exercise 8.4.3
Calculate ΔH for the process:
1
NO(g) + O (g) ⟶ NO (g) ΔH = −57.06 kJ
2 2
2
Answer
66.4 kJ
Here is a less straightforward example that illustrates the thought process involved in solving many Hess’s law problems. It shows
how we can find many standard enthalpies of formation (and other values of ΔH) if they are difficult to determine experimentally.
8.4.4 https://chem.libretexts.org/@go/page/222609
Using Hess’s Law Chlorine monofluoride can react with fluorine to form chlorine trifluoride:
(i) ClF(g) + F 2
(g) ⟶ ClF (g)
3
ΔH ° = ?
Use the reactions here to determine the ΔH° for reaction (i):
(ii) 2 OF 2
(g) ⟶ O (g) + 2 F (g)
2 2
ΔH
∘
(ii)
= −49.4 kJ
(iii) 2 ClF(g) + O 2
(g) ⟶ Cl O(g) + OF (g)
2 2
ΔH
∘
(iii)
= +205.6 kJ
3
(iv) ClF 3
(g) + O (g) ⟶
2
1
2
Cl O(g) +
2
OF (g)
2
ΔH
∘
(iv)
= +266.7 kJ
2
Solution
Our goal is to manipulate and combine reactions (ii), (iii), and (iv) such that they add up to reaction (i). Going from left to right
in (i), we first see that ClF is needed as a reactant. This can be obtained by multiplying reaction (iii) by , which means that
(g)
1
1 1 1 1
ClF(g) + O (g) ⟶ Cl O(g) + OF (g) ΔH ° = (205.6) = +102.8 kJ
2 2 2
2 2 2 2
Next, we see that F is also needed as a reactant. To get this, reverse and halve reaction (ii), which means that the ΔH° changes
2
Now check to make sure that these reactions add up to the reaction we want:
1 1 1
ClF(g) + O (g) ⟶ Cl O(g) + OF (g) ΔH ° = +102.8 kJ
2 2 2
2 2 2
1
O (g) + F (g) ⟶ OF (g) ΔH ° = +24.7 kJ
2 2 2
2
1 3
Cl O(g) + OF (g) ⟶ ClF (g) + O (g) ΔH ° = −266.7 kJ
2 2 3 2
2 2
¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯
¯ ¯
¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯¯
¯
ClF(g) + F ⟶ ClF (g) ΔH ° = −139.2 kJ
2 3
3
Reactants 1
2
O
2
and 1
2
O
2
cancel out product O2; product 1
2
Cl O
2
cancels reactant 1
2
Cl O
2
; and reactant OF
2
is cancelled by
2
products OF and OF2. This leaves only reactants ClF(g) and F2(g) and product ClF3(g), which are what we want. Since
1
2 2
summing these three modified reactions yields the reaction of interest, summing the three modified ΔH° values will give the
desired ΔH°:
ΔH ° = (+102.8 kJ) + (24.7 kJ) + (−266.7 kJ) = −139.2 kJ
Exercise 8.4.4
Aluminum chloride can be formed from its elements:
(i) 2 Al(s) + 3 Cl 2
(g) ⟶ 2 AlCl (s)
3
ΔH ° = ?
Use the reactions here to determine the ΔH° for reaction (i):
(ii) HCl(g) ⟶ HCl(aq) ΔH
(ii)
∘
= −74.8 kJ
(iii) H 2
(g) + Cl (g) ⟶ 2 HCl(g)
2
ΔH
∘
(iii)
= −185 kJ
(iv) AlCl 3
(aq) ⟶ AlCl (s)
3
ΔH
∘
(iv)
= +323 kJ/mol
8.4.5 https://chem.libretexts.org/@go/page/222609
Answer
−1407 kJ
We also can use Hess’s law to determine the enthalpy change of any reaction if the corresponding enthalpies of formation of the
reactants and products are available. The stepwise reactions we consider are: (i) decompositions of the reactants into their
component elements (for which the enthalpy changes are proportional to the negative of the enthalpies of formation of the
reactants), followed by (ii) re-combinations of the elements to give the products (with the enthalpy changes proportional to the
enthalpies of formation of the products). The standard enthalpy change of the overall reaction is therefore equal to: (ii) the sum of
the standard enthalpies of formation of all the products plus (i) the sum of the negatives of the standard enthalpies of formation of
the reactants. This is usually rearranged slightly to be written as follows, with ∑ representing “the sum of” and n standing for the
stoichiometric coefficients:
∘ ∘ ∘
ΔH = ∑ n × ΔH (products) − ∑ n × ΔH (reactants) (8.4.11)
reaction f f
The following example shows in detail why this equation is valid, and how to use it to calculate the enthalpy change for a reaction
of interest.
⎡ −207.4 kJ +90.2 kJ ⎤
= 2 mol HNO3 × +1 mol NO (g) ×
⎣ mol HNO3 (aq) mol NO (g) ⎦
⎡ +33.2 kJ −285.8 kJ ⎤
− 3 mol NO2 (g) × +1 mol H2 O (l) ×
⎣ mol NO2 (g) mol H2 O (l) ⎦
= −138.4 kJ
3
∘
3 NO (g) ⟶ N (g) + 3 O (g) ΔH = −99.6 kJ
2 2 2 1
2
1
∘ ∘
H O(l) ⟶ H (g) + O (g) ΔH = +285.8 kJ [−1 × ΔH (H O)]
2 2 2 2 f 2
2
∘ ∘
H (g) + N (g) + 3 O (g) ⟶ 2 HNO (aq) ΔH = −414.8 kJ [2 × ΔH (HNO )]
2 2 2 3 3 f 3
1 1
∘
N (g) + O (g) ⟶ NO(g) ΔH = +90.2 kJ [1 × (NO)]
2 2 4
2 2
Summing these reaction equations gives the reaction we are interested in:
8.4.6 https://chem.libretexts.org/@go/page/222609
∘ ∘ ∘ ∘ ∘
ΔHrxn = ΔH + ΔH + ΔH + ΔH = (−99.6 kJ) + (+285.8 kJ) + (−414.8 kJ) + (+90.2 kJ)
1 2 3 4
= −138.4 kJ
So the standard enthalpy change for this reaction is ΔH° = −138.4 kJ.
Note that this result was obtained by:
1. multiplying the ΔH of each product by its stoichiometric coefficient and summing those values,
∘
f
2. multiplying the ΔH of each reactant by its stoichiometric coefficient and summing those values, and then
∘
f
3. subtracting the result found in step 2 from the result found in step 1.
This is also the procedure in using the general equation, as shown.
Exercise 8.4.5
Calculate the heat of combustion of 1 mole of ethanol, C2H5OH(l), when H2O(l) and CO2(g) are formed. Use the following
enthalpies of formation: C2H5OH(l), −278 kJ/mol; H2O(l), −286 kJ/mol; and CO2(g), −394 kJ/mol.
Answer
−1368 kJ/mol
Summary
Video 8.4.1 : A summary of Enthalpy and Hess' Law from Crash Course Chemistry.
If a chemical change is carried out at constant pressure and the only work done is caused by expansion or contraction, q for the
change is called the enthalpy change with the symbol ΔH, or ΔH for reactions occurring under standard state conditions. The
∘
298
value of ΔH for a reaction in one direction is equal in magnitude, but opposite in sign, to ΔH for the reaction in the opposite
direction, and ΔH is directly proportional to the quantity of reactants and products. Examples of enthalpy changes include enthalpy
of combustion, enthalpy of fusion, enthalpy of vaporization, and standard enthalpy of formation. The standard enthalpy of
formation, ΔH , is the enthalpy change accompanying the formation of 1 mole of a substance from the elements in their most
∘
f
stable states at 1 bar (standard state). Many of the processes are carried out at 298.15 K. If the enthalpies of formation are available
for the reactants and products of a reaction, the enthalpy change can be calculated using Hess’s law: If a process can be written as
the sum of several stepwise processes, the enthalpy change of the total process equals the sum of the enthalpy changes of the
various steps.
Key Equations
ΔU = q + w
∘ ∘ ∘
ΔH = ∑ n × ΔH (products) − ∑ n × ΔH (reactants)
reaction f f
Footnotes
1. For more on algal fuel, see http://www.theguardian.com/environme...n-fuel-problem.
8.4.7 https://chem.libretexts.org/@go/page/222609
Glossary
chemical thermodynamics
area of science that deals with the relationships between heat, work, and all forms of energy associated with chemical and
physical processes
enthalpy (H)
sum of a system’s internal energy and the mathematical product of its pressure and volume
Hess’s law
if a process can be represented as the sum of several steps, the enthalpy change of the process equals the sum of the enthalpy
changes of the steps
hydrocarbon
compound composed only of hydrogen and carbon; the major component of fossil fuels
heat released when one mole of a compound undergoes complete combustion under standard conditions
enthalpy change of a chemical reaction in which 1 mole of a pure substance is formed from its elements in their most stable
states under standard state conditions
standard state
set of physical conditions as accepted as common reference conditions for reporting thermodynamic properties; 1 bar of
pressure, and solutions at 1 molar concentrations, usually at a temperature of 298.15 K
state function
property depending only on the state of a system, and not the path taken to reach that state
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide E. Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
8.4: Standard Enthalpy and Hess’ Law is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
8.4.8 https://chem.libretexts.org/@go/page/222609
8.4.1: Practice Problems- Enthalpy and Hess’ Law
PROBLEM 8.4.1.1
How much heat is produced by burning 4.00 moles of acetylene under standard state conditions?
Answer
5204.4 kJ
PROBLEM 8.4.1.2
How much heat is produced by combustion of 125 g of methanol (CH3OH) under standard state conditions?
Answer
2836.3 kJ
Problem 8.2.2
PROBLEM 8.4.1.3
How many moles of isooctane must be burned to produce 100 kJ of heat under standard state conditions?
Answer
1.83 × 10−2 mol
PROBLEM 8.4.1.4
What mass of carbon monoxide must be burned to produce 175 kJ of heat under standard state conditions?
Answer
17.3 g
8.4.1.1 https://chem.libretexts.org/@go/page/217302
Problem 8.2.4
PROBLEM 8.4.1.5
When 2.50 g of methane burns in oxygen, 125 kJ of heat is produced. What is the enthalpy of combustion per mole of methane
under these conditions?
Answer
802 kJ mol−1
PROBLEM 8.4.1.6
a. How much heat is produced when 100 mL of 0.250 M HCl (density, 1.00 g/mL) and 200 mL of 0.150 M NaOH (density, 1.00
g/mL) are mixed? (Refer to Example 8.2.1 for how the heat of this reaction was derived)
∘
HCl(aq) + NaOH(aq) ⟶ NaCl(aq) + H O(l) ΔH = −58 kJ/mol (8.4.1.1)
2 298
b. If both solutions are at the same temperature and the heat capacity of the products is 4.19 J/g °C, how much will the
temperature increase? What assumption did you make in your calculation?
Answer a
-1.45 kJ
Answer b
-1.15 °C, assuming the products have the same density as the reactants (1.00 g/mL)
Problem 8.2.6
8.4.1.2 https://chem.libretexts.org/@go/page/217302
PROBLEM 8.4.1.7
A sample of 0.562 g of carbon is burned in oxygen in a bomb calorimeter, producing carbon dioxide. Assume both the reactants
and products are under standard state conditions, and that the heat released is directly proportional to the enthalpy of combustion
of graphite. The temperature of the calorimeter increases from 26.74 °C to 27.93 °C. What is the heat capacity of the calorimeter
and its contents?
Answer
15.5 kJ/ºC
PROBLEM 8.4.1.8
Homes may be heated by pumping hot water through radiators. What mass of water will provide the same amount of heat when
cooled from 95.0 to 35.0 °C, as the heat provided when 100 g of steam is cooled from 110 °C to 100 °C.
Answer
7.43 g
PROBLEM 8.4.1.9
The following sequence of reactions occurs in the commercial production of aqueous nitric acid:
4 NH (g) + 5 O (g) ⟶ 4 NO(g) + 6 H O(l) ΔH = −907 kJ
3 2 2
Determine the total energy change for the production of one mole of aqueous nitric acid by this process.
Answer
495 kJ/mol
PROBLEM 8.4.1.10
Both graphite and diamond burn.
C(s, diamond) + O (g) ⟶ CO (g)
2 2
Which produces more heat, the combustion of graphite or the combustion of diamond? (Hint: The heats of formation for all
these compounds can be found in Table T1.
Answer
Diamond
PROBLEM 8.4.1.11
Calculate ΔH for the process
Hg Cl (s) ⟶ 2 Hg(l) + Cl (g)
2 2 2
8.4.1.3 https://chem.libretexts.org/@go/page/217302
Answer
265 kJ
PROBLEM 8.4.1.12
Using the data in Table T1, calculate the standard enthalpy change for each of the following reactions:
a. N (g) + O (g) ⟶ 2 NO(g)
2 2
d. 2 LiOH(s) + CO (g) ⟶ Li CO (s) + H O(g) (Hint: For LiOH(s), ΔHf = -487.5 kJ/mol; For Li2CO3 (s) , ΔHf =
2 2 3 2
-1216.04 kJ/mol)
Answer
182.6 kJ
Answer
-657.0 kJ mol-1
Answer
-33.2 kJ
Answer
-89.34 kJ
Problem 8.2.12
PROBLEM 8.4.1.13
Using the data in Table T1, calculate the standard enthalpy change for each of the following reactions:
a. Si(s) + 2 F (g) ⟶ SiF (g) (Hint: For SiF4 (g), ΔHf = -1615.0 kJ/mol)
2 4
b. 2 C(s) + 2 H (g) + O (g) ⟶ CH CO H(l) (Hint: For CH3CO2H(l), ΔHf = -484.3 kJ/mol)
2 2 3 2
d. CS (g) + 3 Cl (g) ⟶ CCl (g) + S Cl (g) (Hint: For S2Cl2(g), ΔHf = -19.5 kJ/mol)
2 2 4 2 2
Answer a
−1615.0 kJ mol−1
Answer b
8.4.1.4 https://chem.libretexts.org/@go/page/217302
−484.3 kJ mol−1
Answer c
164.2 kJ
Answer d
−232.1 kJ
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
8.4.1: Practice Problems- Enthalpy and Hess’ Law is shared under a not declared license and was authored, remixed, and/or curated by
LibreTexts.
8.4.1.5 https://chem.libretexts.org/@go/page/217302
8.5: Calorimetry
Skills to Develop
Explain the technique of calorimetry
Calculate and interpret heat and related properties using typical calorimetry data
One technique we can use to measure the amount of heat involved in a chemical or physical process is known as calorimetry. Calorimetry is used
to measure amounts of heat transferred to or from a substance. To do so, the heat is exchanged with a calibrated object (calorimeter). The change
in temperature of the measuring part of the calorimeter is converted into the amount of heat (since the previous calibration was used to establish
its heat capacity). The measurement of heat transfer using this approach requires the definition of a system (the substance or substances
undergoing the chemical or physical change) and its surroundings (the other components of the measurement apparatus that serve to either
provide heat to the system or absorb heat from the system). Knowledge of the heat capacity of the surroundings, and careful measurements of the
masses of the system and surroundings and their temperatures before and after the process allows one to calculate the heat transferred as
described in this section.
A calorimeter is a device used to measure the amount of heat involved in a chemical or physical process. For example, when an exothermic
reaction occurs in solution in a calorimeter, the heat produced by the reaction is absorbed by the solution, which increases its temperature. When
an endothermic reaction occurs, the heat required is absorbed from the thermal energy of the solution, which decreases its temperature (Figure
8.5.1). The temperature change, along with the specific heat and mass of the solution, can then be used to calculate the amount of heat involved
in either case.
Figure 8.5.1 : In a calorimetric determination, either (a) an exothermic process occurs and heat, q, is negative, indicating that thermal energy is
transferred from the system to its surroundings, or (b) an endothermic process occurs and heat, q, is positive, indicating that thermal energy is
transferred from the surroundings to the system.
By convention, q is given a negative (-) sign when the system releases heat to the surroundings (exothermic); q is given a positive (+) sign
when the system absorbs heat from the surroundings (endothermic).
Scientists use well-insulated calorimeters that all but prevent the transfer of heat between the calorimeter and its environment. This enables the
accurate determination of the heat involved in chemical processes, the energy content of foods, and so on. General chemistry students often use
simple calorimeters constructed from polystyrene cups (Figure 8.5.2). These easy-to-use “coffee cup” calorimeters allow more heat exchange
with their surroundings, and therefore produce less accurate energy values.
8.5.1 https://chem.libretexts.org/@go/page/217299
Figure 8.5.2 : A simple calorimeter can be constructed from two polystyrene cups. A thermometer and stirrer extend through the cover into the
reaction mixture.
Commercial solution calorimeters are also available. Relatively inexpensive calorimeters often consist of two thin-walled cups that are nested in a
way that minimizes thermal contact during use, along with an insulated cover, handheld stirrer, and simple thermometer. More expensive
calorimeters used for industry and research typically have a well-insulated, fully enclosed reaction vessel, motorized stirring mechanism, and a
more accurate temperature sensor (Figure 8.5.3).
Figure 8.5.3 : Commercial solution calorimeters range from (a) simple, inexpensive models for student use to (b) expensive, more accurate
models for industry and research.
Before we practice calorimetry problems involving chemical reactions, consider a simple example that illustrates the core idea behind
calorimetry. Suppose we initially have a high-temperature substance, such as a hot piece of metal (M), and a low-temperature substance, such as
cool water (W). If we place the metal in the water, heat will flow from M to W. The temperature of M will decrease, and the temperature of W
will increase, until the two substances have the same temperature—that is, when they reach thermal equilibrium (Figure 8.5.4). If this occurs in a
calorimeter, ideally all of this heat transfer occurs between the two substances, with no heat gained or lost by either the calorimeter or the
calorimeter’s surroundings. Under these ideal circumstances, the net heat change is zero:
This relationship can be rearranged to show that the heat gained by substance M is equal to the heat lost by substance W:
The magnitude of the heat (change) is therefore the same for both substances, and the negative sign merely shows that qsubstance M and qsubstance
W are opposite in direction of heat flow (gain or loss) but does not indicate the arithmetic sign of either q value (that is determined by whether the
matter in question gains or loses heat, per definition). In the specific situation described, qsubstance M is a negative value and qsubstance W is
positive, since heat is transferred from M to W.
8.5.2 https://chem.libretexts.org/@go/page/217299
Figure 8.5.4 : In a simple calorimetry process, (a) heat, q, is transferred from the hot metal, M, to the cool water, W, until (b) both are at the
same temperature.
crebar × mrebar × (Tf ,rebar − Ti,rebar) = −cwater × mwater × (Tf ,water − Ti,water) (8.5.5)
The density of water is 1.0 g/mL, so 425 mL of water = 425 g. Noting that the final temperature of both the rebar and water is 42.7 °C,
substituting known values yields:
(0.449 J/g °C)(360g)(42.7°C − Ti,rebar) = −(4.184 J/g °C)(425 g)(42.7°C − 24.0°C) (8.5.6)
Solving this gives Ti,rebar= 248 °C, so the initial temperature of the rebar was 248 °C.
Exercise 8.5.1A
A 248-g piece of copper is dropped into 390 mL of water at 22.6 °C. The final temperature of the water was measured as 39.9 °C. Calculate
the initial temperature of the piece of copper. Assume that all heat transfer occurs between the copper and the water.
Answer
The initial temperature of the copper was 335.6 °C.
Exercise 8.5.1B
A 248-g piece of copper initially at 314 °C is dropped into 390 mL of water initially at 22.6 °C. Assuming that all heat transfer occurs between
the copper and the water, calculate the final temperature.
Answer
8.5.3 https://chem.libretexts.org/@go/page/217299
The final temperature (reached by both copper and water) is 38.7 °C.
This method can also be used to determine other quantities, such as the specific heat of an unknown metal.
Noting that since the metal was submerged in boiling water, its initial temperature was 100.0 °C; and that for water, 60.0 mL = 60.0 g; we
have:
(cmetal )(59.7 g)(28.5°C − 100.0°C) = −(4.18 J/g °C)(60.0 g)(28.5°C − 22.0°C) (8.5.10)
Solving this:
−(4.184 J/g °C)(60.0 g)(6.5°C)
cmetal = = 0.38 J/g °C (8.5.11)
(59.7 g)(−71.5°C)
Comparing this with values in Table T4, our experimental specific heat is closest to the value for copper (0.39 J/g °C), so we identify the metal
as copper.
Exercise 8.5.2
A 92.9-g piece of a silver/gray metal is heated to 178.0 °C, and then quickly transferred into 75.0 mL of water initially at 24.0 °C. After 5
minutes, both the metal and the water have reached the same temperature: 29.7 °C. Determine the specific heat and the identity of the metal.
(Note: You should find that the specific heat is close to that of two different metals. Explain how you can confidently determine the identity of
the metal).
Answer
cmetal = 0.13 J/g °C
This specific heat is close to that of either gold or lead. It would be difficult to determine which metal this was based solely on the
numerical values. However, the observation that the metal is silver/gray in addition to the value for the specific heat indicates that the
metal is lead.
When we use calorimetry to determine the heat involved in a chemical reaction, the same principles we have been discussing apply. The amount
of heat absorbed by the calorimeter is often small enough that we can neglect it (though not for highly accurate measurements, as discussed later),
and the calorimeter minimizes energy exchange with the surroundings. Because energy is neither created nor destroyed during a chemical
reaction, there is no overall energy change during the reaction. The heat produced or consumed in the reaction (the “system”), qreaction, plus the
heat absorbed or lost by the solution (the “surroundings”), qsolution, must add up to zero:
qreaction + qsolution = 0 (8.5.12)
This means that the amount of heat produced or consumed in the reaction equals the amount of heat absorbed or lost by the solution:
qreaction = −qsolution (8.5.13)
This concept lies at the heart of all calorimetry problems and calculations.
Solution
8.5.4 https://chem.libretexts.org/@go/page/217299
To visualize what is going on, imagine that you could combine the two solutions so quickly that no reaction took place while they mixed; then
after mixing, the reaction took place. At the instant of mixing, you have 100.0 mL of a mixture of HCl and NaOH at 22.0 °C. The HCl and
NaOH then react until the solution temperature reaches 28.9 °C.
The heat given off by the reaction is equal to that taken in by the solution. Therefore:
qreaction = −qsolution
(It is important to remember that this relationship only holds if the calorimeter does not absorb any heat from the reaction, and there is no heat
exchange between the calorimeter and its surroundings.)
Next, we know that the heat absorbed by the solution depends on its specific heat, mass, and temperature change:
qsolution = (c × m × ΔT )solution
To proceed with this calculation, we need to make a few more reasonable assumptions or approximations. Since the solution is aqueous, we
can proceed as if it were water in terms of its specific heat and mass values. The density of water is approximately 1.0 g/mL, so 100.0 mL has
a mass of about 1.0 × 102 g (two significant figures). The specific heat of water is approximately 4.18 J/g °C, so we use that for the specific
heat of the solution. Substituting these values gives:
2 3
qsolution = (4.184 J/g °C)(1.0 × 10 g)(28.9°C − 22.0°C) = 2.89 × 10 J (8.5.14)
Finally, since we are trying to find the heat of the reaction, we have:
3
qreaction = −qsolution = −2.89 × 10 J
The negative sign indicates that the reaction is exothermic. It produces 2.89 kJ of heat.
Exercise 8.5.3
When 100 mL of 0.200 M NaCl(aq) and 100 mL of 0.200 M AgNO3(aq), both at 21.9 °C, are mixed in a coffee cup calorimeter, the
temperature increases to 23.5 °C as solid AgCl forms. How much heat is produced by this precipitation reaction? What assumptions did you
make to determine your value?
Answer
J; assume no heat is absorbed by the calorimeter, no heat is exchanged between the calorimeter and its surroundings, and that
3
1.34 × 10
the specific heat and mass of the solution are the same as those for water
warming them (at least for a while). If the hand warmer is reheated, the NaC2H3O2 redissolves and can be reused.
Figure 8.5.5 : Chemical hand warmers produce heat that warms your hand on a cold day. In this one, you can see the metal disc that initiates
the exothermic precipitation reaction. (credit: modification of work by Science Buddies TV/YouTube)
Another common hand warmer produces heat when it is ripped open, exposing iron and water in the hand warmer to oxygen in the air. One
simplified version of this exothermic reaction is
3
2 Fe(s) + O (g) ⟶ Fe O (s). (8.5.15)
2 2 3
2
Salt in the hand warmer catalyzes the reaction, so it produces heat more rapidly; cellulose, vermiculite, and activated carbon help distribute the
heat evenly. Other types of hand warmers use lighter fluid (a platinum catalyst helps lighter fluid oxidize exothermically), charcoal (charcoal
oxidizes in a special case), or electrical units that produce heat by passing an electrical current from a battery through resistive wires.
8.5.5 https://chem.libretexts.org/@go/page/217299
Example 8.5.4 : Heat Flow in an Instant Ice Pack
When solid ammonium nitrate dissolves in water, the solution becomes cold. This is the basis for an “instant ice pack” (Figure 8.5.5). When
3.21 g of solid NH4NO3 dissolves in 50.0 g of water at 24.9 °C in a calorimeter, the temperature decreases to 20.3 °C.
Calculate the value of q for this reaction and explain the meaning of its arithmetic sign. State any assumptions that you made.
Figure 8.5.5 : An instant cold pack consists of a bag containing solid ammonium nitrate and a second bag of water. When the bag of water is
broken, the pack becomes cold because the dissolution of ammonium nitrate is an endothermic process that removes thermal energy from the
water. The cold pack then removes thermal energy from your body.
Solution
We assume that the calorimeter prevents heat transfer between the solution and its external environment (including the calorimeter itself), in
which case:
qrxn = −qsoln (8.5.16)
with “rxn” and “soln” used as shorthand for “reaction” and “solution,” respectively.
Assuming also that the specific heat of the solution is the same as that for water, we have:
qrxn = −qsoln = −(c × m × ΔT )soln
3
+ 1.0 × 10 J = +1.0 kJ
The positive sign for q indicates that the dissolution is an endothermic process.
Exercise 8.5.4
When a 3.00-g sample of KCl was added to 3.00 × 102 g of water in a coffee cup calorimeter, the temperature decreased by 1.05 °C. How
much heat is involved in the dissolution of the KCl? What assumptions did you make?
Answer
1.33 kJ; assume that the calorimeter prevents heat transfer between the solution and its external environment (including the calorimeter
itself) and that the specific heat of the solution is the same as that for water.
If the amount of heat absorbed by a calorimeter is too large to neglect or if we require more accurate results, then we must take into account the
heat absorbed both by the solution and by the calorimeter.
Figure 8.5.6 : (a) A bomb calorimeter is used to measure heat produced by reactions involving gaseous reactants or products, such as
combustion. (b) The reactants are contained in the gas-tight “bomb,” which is submerged in water and surrounded by insulating materials.
8.5.6 https://chem.libretexts.org/@go/page/217299
(credit a: modification of work by “Harbor1”/Wikimedia commons)
The calorimeters described are designed to operate at constant (atmospheric) pressure and are convenient to measure heat flow accompanying
processes that occur in solution. A different type of calorimeter that operates at constant volume, colloquially known as a bomb calorimeter, is
used to measure the energy produced by reactions that yield large amounts of heat and gaseous products, such as combustion reactions. (The term
“bomb” comes from the observation that these reactions can be vigorous enough to resemble explosions that would damage other calorimeters.)
This type of calorimeter consists of a robust steel container (the “bomb”) that contains the reactants and is itself submerged in water (Figure
8.5.6). The sample is placed in the bomb, which is then filled with oxygen at high pressure. A small electrical spark is used to ignite the sample.
The energy produced by the reaction is trapped in the steel bomb and the surrounding water. The temperature increase is measured and, along
with the known heat capacity of the calorimeter, is used to calculate the energy produced by the reaction. Bomb calorimeters require calibration
to determine the heat capacity of the calorimeter and ensure accurate results. The calibration is accomplished using a reaction with a known q,
such as a measured quantity of benzoic acid ignited by a spark from a nickel fuse wire that is weighed before and after the reaction. The
temperature change produced by the known reaction is used to determine the heat capacity of the calorimeter. The calibration is generally
performed each time before the calorimeter is used to gather research data.
Video 8.5.1 : Video of view how a bomb calorimeter is prepared for action.
= −[(4.184 J/g °C) × (775 g) × (35.6°C − 23.8°C) + 893 J/°C × (35.6°C − 23.8°C)]
This reaction released 48.7 kJ of heat when 3.12 g of glucose was burned.
Exercise 8.5.5
When 0.963 g of benzene, C6H6, is burned in a bomb calorimeter, the temperature of the calorimeter increases by 8.39 °C. The bomb has a
heat capacity of 784 J/°C and is submerged in 925 mL of water. How much heat was produced by the combustion of the glucose sample?
8.5.7 https://chem.libretexts.org/@go/page/217299
Answer
39.0 kJ
Since the first one was constructed in 1899, 35 calorimeters have been built to measure the heat produced by a living person.1 These whole-body
calorimeters of various designs are large enough to hold an individual human being. More recently, whole-room calorimeters allow for relatively
normal activities to be performed, and these calorimeters generate data that more closely reflect the real world. These calorimeters are used to
measure the metabolism of individuals under different environmental conditions, different dietary regimes, and with different health conditions,
such as diabetes. In humans, metabolism is typically measured in Calories per day. A nutritional calorie (Calorie) is the energy unit used to
quantify the amount of energy derived from the metabolism of foods; one Calorie is equal to 1000 calories (1 kcal), the amount of energy needed
to heat 1 kg of water by 1 °C.
Measuring Nutritional Calories
In your day-to-day life, you may be more familiar with energy being given in Calories, or nutritional calories, which are used to quantify the
amount of energy in foods. One calorie (cal) = exactly 4.184 joules, and one Calorie (note the capitalization) = 1000 cal, or 1 kcal. (This is
approximately the amount of energy needed to heat 1 kg of water by 1 °C.)
The macronutrients in food are proteins, carbohydrates, and fats or oils. Proteins provide about 4 Calories per gram, carbohydrates also
provide about 4 Calories per gram, and fats and oils provide about 9 Calories/g. Nutritional labels on food packages show the caloric content
of one serving of the food, as well as the breakdown into Calories from each of the three macronutrients (Figure 8.5.7).
Figure 8.5.7 : (a) Macaroni and cheese contain energy in the form of the macronutrients in the food. (b) The food’s nutritional information is
shown on the package label. In the US, the energy content is given in Calories (per serving); the rest of the world usually uses kilojoules.
(credit a: modification of work by “Rex Roof”/Flickr)
For the example shown in (b), the total energy per 228-g portion is calculated by:
(5 g protein × 4 Calories/g) + (31 g carb × 4 Calories/g) + (12 g f at × 9 Calories/g) = 252 Calories (8.5.17)
So, you can use food labels to count your Calories. But where do the values come from? And how accurate are they? The caloric content of
foods can be determined by using bomb calorimetry; that is, by burning the food and measuring the energy it contains. A sample of food is
weighed, mixed in a blender, freeze-dried, ground into powder, and formed into a pellet. The pellet is burned inside a bomb calorimeter, and
the measured temperature change is converted into energy per gram of food.
Today, the caloric content on food labels is derived using a method called the Atwater system that uses the average caloric content of the
different chemical constituents of food, protein, carbohydrate, and fats. The average amounts are those given in the equation and are derived
from the various results given by bomb calorimetry of whole foods. The carbohydrate amount is discounted a certain amount for the fiber
content, which is indigestible carbohydrate. To determine the energy content of a food, the quantities of carbohydrate, protein, and fat are each
multiplied by the average Calories per gram for each and the products summed to obtain the total energy.
Summary
8.5.8 https://chem.libretexts.org/@go/page/217299
Calorimetry: Crash Course Chemistry #19
Footnotes
1. 1 Francis D. Reardon et al. “The Snellen human calorimeter revisited, re-engineered and upgraded: Design and performance characteristics.”
Medical and Biological Engineering and Computing 8 (2006)721–28, http://link.springer.com/article/10....517-006-0086-5.
Glossary
bomb calorimeter
device designed to measure the energy change for processes occurring under conditions of constant volume; commonly used for reactions
involving solid and gaseous reactants or products
calorimeter
device used to measure the amount of heat absorbed or released in a chemical or physical process
calorimetry
process of measuring the amount of heat involved in a chemical or physical process
surroundings
all matter other than the system being studied
system
portion of matter undergoing a chemical or physical change being studied
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/[email protected]).
Adelaide E. Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
8.5: Calorimetry is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
8.5.9 https://chem.libretexts.org/@go/page/217299
8.5.1: Practice Problems- Calorimetry
PROBLEM 8.5.1.1
A 500-mL bottle of water at room temperature and a 2-L bottle of water at the same temperature were placed in a refrigerator.
After 30 minutes, the 500-mL bottle of water had cooled to the temperature of the refrigerator. An hour later, the 2-L of water
had cooled to the same temperature. When asked which sample of water lost the most heat, Student A replied that both bottles
lost the same amount of heat because they started at the same temperature and finished at the same temperature. Student B
thought that the 2-L bottle of water lost more heat because there was more water. A third student believed that the 500-mL bottle
of water lost more heat because it cooled more quickly. A fourth student thought that it was not possible to tell because we do
not know the initial temperature and the final temperature of the water. Indicate which of these answers is correct and describe
the error in each of the other answers.
Answer
Student A is incorrect because the mass of water in both containers is not the same.
Student C is incorrect because the bottle cooled quicker due to less mass of water.
Student D is incorrect because it doesn't matter what the change in temperature is as long as it is the same for both bottles.
Student B is correct: if the change in temperature is the same, the one with the more mass (the 2L bottle) had more heat loss.
We could prove this using q = c × m × ΔT = c × m × (T −T
final )
initial from Section 8.1.
PROBLEM 8.5.1.2
How many milliliters of water at 23 °C with a density of 1.00 g/mL must be mixed with 180 mL (about 6 oz) of coffee at 95 °C
so that the resulting combination will have a temperature of 60 °C? Assume that coffee and water have the same density and the
same specific heat (4.184 J/g °C).
Answer
170 mL
Problem 8.2.2
PROBLEM 8.5.1.3
How much will the temperature of a cup (180 g) of coffee at 95 °C be reduced when a 45 g silver spoon (specific heat 0.24 J/g
°C) at 25 °C is placed in the coffee and the two are allowed to reach the same temperature? Assume that the coffee has the same
density and specific heat as water.
8.5.1.1 https://chem.libretexts.org/@go/page/217300
Answer
The temperature of the coffee will drop 1 degree.
PROBLEM 8.5.1.4
A 45-g aluminum spoon (specific heat 0.88 J/g °C) at 24 °C is placed in 180 mL (180 g) of coffee at 85 °C and the temperature
of the two become equal.
a. What is the final temperature when the two become equal? Assume that coffee has the same specific heat as water.
b. The first time a student solved this problem she got an answer of 88 °C. Explain why this is clearly an incorrect answer.
Answer a
81.95 °C
Answer b
This temperature is higher than the starting temperature of the coffee, which is impossible.
Problem 8.2.4
PROBLEM 8.5.1.5
The temperature of the cooling water as it leaves the hot engine of an automobile is 240 °F. After it passes through the radiator it
has a temperature of 175 °F. Calculate the amount of heat transferred from the engine to the surroundings by one gallon of water
with a specific heat of 4.184 J/g °C.
Answer
2
5.7 × 10 kJ
PROBLEM 8.5.1.6
When 50.0 g of 0.200 M NaCl(aq) at 24.1 °C is added to 100.0 g of 0.100 M AgNO3(aq) at 24.1 °C in a calorimeter, the
temperature increases to 25.2 °C as AgCl(s) forms. Assuming the specific heat of the solution and products is 4.20 J/g °C,
calculate the approximate amount of heat in joules produced.
Answer
693 J
8.5.1.2 https://chem.libretexts.org/@go/page/217300
Problem 8.2.6
PROBLEM 8.5.1.7
The addition of 3.15 g of Ba(OH)2•8H2O to a solution of 1.52 g of NH4SCN in 100 g of water in a calorimeter caused the
temperature to fall by 3.1 °C. Assuming the specific heat of the solution and products is 4.20 J/g °C, calculate the approximate
amount of heat absorbed by the reaction, which can be represented by the following equation:
Ba(OH )2 ⋅ 8 H2 O(s) + 2N H4 SC N(aq) → Ba(SC N )2(aq) + 2N H3(aq) + 10 H2 O(l) (8.5.1.1)
Answer
1.4 kJ
PROBLEM 8.5.1.8
When 1.0 g of fructose, C6H12O6(s), a sugar commonly found in fruits, is burned in oxygen in a bomb calorimeter, the
temperature of the calorimeter increases by 1.58 °C. If the heat capacity of the calorimeter and its contents is 9.90 kJ/°C, what is
q for this combustion?
Answer
15.64 kJ
Problem 8.2.8
PROBLEM 8.5.1.9
8.5.1.3 https://chem.libretexts.org/@go/page/217300
One method of generating electricity is by burning coal to heat water, which produces steam that drives an electric generator. To
determine the rate at which coal is to be fed into the burner in this type of plant, the heat of combustion per ton of coal must be
determined using a bomb calorimeter. When 1.00 g of coal is burned in a bomb calorimeter, the temperature increases by 1.48
°C. If the heat capacity of the calorimeter is 21.6 kJ/°C, determine the heat produced by combustion of a ton of coal (2000
pounds). Remember 1 kg = 2.2 pounds
Answer
2.91 x 107 kJ
PROBLEM 8.5.1.10
A teaspoon of the carbohydrate sucrose (common sugar) contains 16 Calories (16 kcal). What is the mass of one teaspoon of
sucrose if the average number of Calories for carbohydrates is 4.1 Calories/g?
Answer
3.9 g
Problem 8.2.10
PROBLEM 8.5.1.11
What is the maximum mass of carbohydrate in a 6-oz serving of diet soda that contains less than 1 Calorie per can if the average
number of Calories for carbohydrates is 4.1 Calories/g?
Answer
0.24 g
PROBLEM 8.5.1.12
A pint of premium ice cream can contain 1100 Calories. What mass of fat, in grams and pounds, must be produced in the body
to store an extra 1.1 × 103 Calories if the average number of Calories for fat is 9.1 Calories/g? Remember 1 kg = 2.2 pounds
Answer
120.87 g or 1.2 x 102 g with 2 significant figures
0.266 lbs or 0.27 lbs with 2 significant figures
8.5.1.4 https://chem.libretexts.org/@go/page/217300
Click here to see a video of the solution
Problem 8.2.12
PROBLEM 8.5.1.13
A serving of a breakfast cereal contains 3 g of protein, 18 g of carbohydrates, and 6 g of fat. What is the Calorie content of a
serving of this cereal if the average number of Calories for fat is 9.1 Calories/g, for carbohydrates is 4.1 Calories/g, and for
protein is 4.1 Calories/g?
Answer
1.4 × 102 Calories
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
8.5.1: Practice Problems- Calorimetry is shared under a not declared license and was authored, remixed, and/or curated by LibreTexts.
8.5.1.5 https://chem.libretexts.org/@go/page/217300
Index
S
Significant figures
1.8: Significant Digits
1 https://chem.libretexts.org/@go/page/240545
Glossary
Sample Word 1 | Sample Definition 1
1 https://chem.libretexts.org/@go/page/279469