Introduction To General Chemistry (Muhammad Arif Malik)
Introduction To General Chemistry (Muhammad Arif Malik)
GENERAL CHEMISTRY
(MALIK)
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).
Dedication
Licensing
2: Elements
2.1: Dalton’s atomic theory
2.2: Subatomic particles and a modern view of an atom
2.3: Atoms of elements
2.4: The periodic table
2.5: Electrons in atoms
2.6: The periodic trends in properties of the elements
3: Compounds
3.1: Bonding in compounds
3.2: Naming binary ionic compounds
3.3: Polyatomic ions and their compounds
3.4: Naming acids
3.5: Naming binary covalent compounds
3.6: Lewis structures of molecules
3.7: Molecular shapes –Valence shell electron pair repulsion (VSEPR) theory
3.8: Polarity of molecules
3.9: Intramolecular forces and intermolecular forces
1
4.5: Oxidation-reduction reactions
4.6: Energetics of chemical reactions
4.7: Stoichiometric calculations
5: Solutions
5.1: Introduction to solution
5.2: Solubility
5.3: Electrolytes
5.4: Concentration of solutions
5.5: Osmosis
7: Gases
7.1: Characteristics of gases
7.2: The pressure-volume relationship
7.3: The temperature-volume relationship
7.4: The pressure-temperature relationship
7.5: The combined gas law
7.6: The volume-amount relationship
7.7: Ideal gas law
7.8: Dalton’s law of partial pressure
8: Nuclear chemistry
8.1: Introduction to nuclear chemistry
8.2: Radioactivity
8.3: Half-life of radioisotopes
8.4: Radiation measurements
8.5: Ionizing radiation exposures
8.6: Medical uses of radioisotopes
8.7: Making radioisotopes for medical uses
8.8: Nuclear fusion and fission
Index
Glossary
2
Detailed Licensing
Introduction to General Chemistry (Malik) is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad
Arif Malik.
3
Dedication
Dedicated to Mr. Frank Reidy, a philanthropist who brought the author to the US, and who supported research on the medical
application of pulsed electric fields for medical treatments at Frank Reidy Research Center For Bioelectrics, Old Dominion
University, Norfolk, VA. The research resulted in the development of techniques proven to cure some forms of cancer.
1 https://chem.libretexts.org/@go/page/373514
Licensing
A detailed breakdown of this resource's licensing can be found in Back Matter/Detailed Licensing.
1 https://chem.libretexts.org/@go/page/374366
CHAPTER OVERVIEW
1: Matter energy and their measurements
1.1: Matter and energy
1.2: What is chemistry?
1.3: Measurements
1.4: Significant Figures
1.5: Unit conversions
1.6: Equations and graphs
1.7: Density and specific gravity measurements
1.8: Heat and its measurements
1.9: Heat and changes in physical states of matter
1: Matter energy and their measurements is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif
Malik.
1
1.1: Matter and energy
What is the matter?
The matter is anything that has mass and occupies space.
The matter is a natural material that makes up the universe. The matter is composed of tiny particles called atoms held together by
forces called bonds. The matter is classified as a pure matter if it has a constant and a non-variable composition of the type of
atoms. Pur matter is either an element or a compound.
Element
For example, carbon shown in Fig. 1.1.1 is an element. Elements can not convert to a simpler matter by physical or chemical
methods. There are around a hundred different elements known at this time. For example, hydrogen, oxygen, carbon, nitrogen,
sodium, chlorine, iron, cobalt, gold, and silver are a few elements.
Figure 1.1.1 : The pencil tip is carbon -an element that is composed of one type of atom, as shown in the model.
Symbol of an element
Elements are represented by symbols, the first alphabet of their English or non-English name, written in capital letters. For
example, C for carbon, O for oxygen, and H for hydrogen. Usually, another alphabet is also chosen from the element's name and
written as a small letter, e.g., He for helium, Co for cobalt. Some element symbols are derived from non-English names, e.g., Fe for
iron is from its Latin name Ferrum, and Au for gold is from its Latin name Aurum.
Two consecutive capital letters do not represent an element; they may be two different elements combined.
CO is not a symbol of an element; it is a pure matter which is a combination of carbon, and oxygen, respectively, bonded together
in a 1:1 atom ratio.
Molecule
A molecule is a group of two or more atoms held together by forces called chemical bonds. The molecule is the smallest particle of
matter that can exist freely. A single atom of some elements is capable of existing freely, and it is also considered a molecule. For
example, He, O2, P4, and S8 are examples of elements having molecules composed of one, two, four, and eight atoms of the same
element, respectively, as illustrated in Fig. 1.1.2. Although metal elements exist as a vast number of atoms bonded together by a
special type of bond called metallic bonds, their symbol is that of a single atom, e.g., Fe for iron and Au for gold.
1.1.1 https://chem.libretexts.org/@go/page/371338
Figure 1.1.2 : Models of molecules of some elements.
Compound
A compound is a pure matter composed of atoms of two or more different elements in a constant whole number ratio held
together by chemical bonds.
The symbol of a compound is a combination of its constituent elements with a subscript to the right of the element symbol
representing the whole number ratio of the atoms of the element in the compound. For example, H2O symbolizes a compound
called water composed of hydrogen (H) and oxygen (O) atoms in a 2:1 ratio. Similarly, NaCl is a symbol of a table salt compound
composed of sodium (Na) and chlorine (Cl) atoms in a 1:1 ratio. If the symbol of a compound also represents a molecule of the
compound, it is called a molecular formula. For example, H2O is a molecule formula of water. On the other hand, table salt is
another class of compound composed of a vast number of atoms of its constituent elements arranged in a specific arrangement in
3D space called a crystal lattice, as illustrated in Fig. 1.1.3. When the compound symbol does not represent a molecule, it only
represents the simple whole-number ratio of the constituent elements; it is called the compound's chemical formula. For example,
NaCl is a chemical formula of a compound called table salt.
Figure 1.1.3 : Sodium chloride (NaCl) –a compound, the crystal (left) and model showing crystal-lattice (right). Source:
Halite(Salt)USGOV.jpg and Benjah-bmm27 (talk · contribs) / Public domain.
A chemical compound always contains the number of atoms of its component elements in a fixed ratio, or compounds are
formed by defined mass ratios of reacting elements.
Pure matter, i.e., elements or compounds, has a fixed proportion of atoms of element/s independent of the source or method of their
preparation, is also called a chemical or a substance. The atoms in a compound are held together by attractive forces called
chemical bonds. Constituent elements in a compound can be separated only by breaking chemical bonds and making new chemical
bonds which is a chemical reaction.
Mixture
Two or more pure substances mixed such that ratio of atoms of the constituent elements is variable is called a mixture.
For example, table salt can be mixed with water, but the ratio of salt (NaCl) to water (H2O) can be varied to give less salty or more
salty water in a mixture. A mixture in which the components are thoroughly mixed, and the composition is constant throughout a
given sample is a homogeneous mixture or a solution. For example, table salt dissolved in water is an example of a homogeneous
mixture or a solution. Other examples of homogeneous mixture include sugar dissolved in water; the air, which is a mixture of
nitrogen, oxygen, carbon dioxide, and other gases; and metal alloys like brass, a mix of copper and zinc metals, etc.
1.1.2 https://chem.libretexts.org/@go/page/371338
Figure 1.1.4 : A heterogeneous mixture of iron and sulfur (left) , separated by a magnet (right). Source: Asoult / CC BY
(https://creativecommons.org/licenses/by/4.0)
If the mixture's components are not thoroughly mixed and composition varies within different regions of a given sample, it is a
heterogeneous mixture. For example, sulfur (S) powder mixed with iron (Fe) filling is a heterogeneous mixture shown in Fig.
1.1.4. Other examples of heterogeneous mixtures include smoke which is a mixture of air and carbon particles; smog which is a
mixture of liquid water droplets suspended in air; and orange juice which is a mixture of sugar, water, fiber particles, etc. Fig. 1.1.5
illustrates the classification of the matter described above.
States of matter
Matter exists in one of the four physical states or phases, i.e., solid (s), liquid (l), gas (g), or plasma. Fig. 1.1.6 illustrates the four
states of matter at the molecular level and Fig. 1.1.7 shows examples of the four states of matter.
Figure 1.1.6 : Four states of matter at the molecular level. Source: https://www.hiclipart.com/free-trans...mwpjm/download
1.1.3 https://chem.libretexts.org/@go/page/371338
Figure 1.1.7 : Examples of the four states of matter. Clockwise from the top left are solid, liquid, plasma, and gas, represented by an
ice sculpture, a drop of water, electrical arcing from a tesla coil, and the air around clouds, respectively. Source: Spirit469 / CC BY-
SA (https://creativecommons.org/licenses/by-sa/3.0)
Solid-state
In the solid-state, the particles, i.e., atoms or molecules, are very close to each other and held strongly by intermolecular forces. The
particles can vibrate around their mean positions, but they cannot slide past each other. Expansion and contraction in a solid-state is
negligible. The solid has a fixed shape and a fixed volume.
Liquid state
In the liquid state, the particles are close enough to experience strong intermolecular interactions that usually do not let the particles
cross the liquid boundary, but the particles can move around within the liquid. Consequently, the particles in a liquid can flow and
acquire the shape of the container but have a fixed volume. Expansion and contraction are negligible in the liquid state.
Gas state
In the gas state, the particles are far apart. The intermolecular interactions are negligible in the gas phase due to the large distances
between the particles. The gas molecules move in straight lines in random directions until they collide with other molecules or the
walls of the container. The collisions are elastic; that is, the molecules bounce off like elastic balls, and the total kinetic energy of
the system is conserved. If exposed to space, the particles keep moving into space. In other words, the particles in a gas can flow,
acquire the shape of the container, and expand or contract to fill up the available space. The gases do not have a fixed shape and do
not have a fixed volume.
Plasma state
In the plasma state, the particles are far apart like gases, and a portion of the negative charge of the particles, i.e., electrons, are
separated from the positive charge potion, i.e., the nucleus. In other words, the atoms in the plasma state are ionized. The plasma
state is not common on earth, but it is the universe's most common state of matter. For example, the matter in the sun and stars is in
the plasma state. Examples of the plasma state on earth include the matter in the lightning bolts and electrical sparks.
What is energy
Energy is a quantitative property transferred to an object and recognizable in the form of performing work or as heat or light. In
simple words: energy is the ability to do work.
What we commonly encounter other than matter is energy. There are two basic types of energies, i.e., the kinetic energy of moving
objects and potential energy stored by the position of an object in a force field.
Kinetic energy
The energy of moving objects is kinetic energy. The mathematical form of kinetic energy (KE) is KE = mv , where m is the 1
2
2
mass and v is the velocity of the moving object. Examples of kinetic energy include the energy of all moving objects that we see
around, like moving vehicles or a moving turbine that generates electricity. Thermal energy is also the kinetic energy of the atoms
and molecules in a matter.
Potential energy
Potential energy is due to the position of an object in a force field. Examples of force fields responsible for the potential energy
include electric, magnetic, gravitational, and elastic forces. Examples of potential energies are electrical energy and gravitational
1.1.4 https://chem.libretexts.org/@go/page/371338
energy. Light is potential energy due to moving electric and magnetic fields. Chemical energy is potential energy stored in chemical
bonds in electrostatic potential energy.
Gravitation potential energy is the energy due to position or height relative to earth. The earth attracts other objects with a force
F = mg , where F is the force, m is the object's mass, and g is the acceleration due to gravity. When an object falls, the potential
energy changes to several forms of energy, including kinetic energy, work against friction from air, sound, and work done in
deformations when it hits the ground.
Chemical potential energy is the energy due to chemical bonds that hold the atoms together in a molecule or compound through
electrical forces between negative charge electrons and positive charge nuclei. Bond forming always releases energy, and bond-
breaking absorbs the same energy. Each bond has different bond energy. In chemical reactions, some bonds break, and some bonds
form.
A combination of kinetic and potential energy is also possible, e.g., in mechanical waves. A sound wave is a mechanical wave that
combines kinetic and potential energy –kinetic because particles move and potential due to the elasticity of the material in which
the deformation (sound) is propagating.
Figure 1.1.8 : Combustion of methane –an example of an exothermic reaction that releases energy as the heat used for cooking
(left), and photosynthesis - an example of an endothermic reaction that absorbs energy (right). Source:
https://www.hiclipart.com/free-trans...imhhh/download and https://www.hiclipart.com/free-trans...-clipart-mslhq
The balance between the energy needed in breaking the bonds and the energy released from forming new bonds determines
whether the chemical reaction releases energy or absorbs energy. Chemical reactions that release energy are called exothermic
reactions. For example, the combustion of methane gas in kitchen burners releases energy. A reaction that absorbs energy is called
an endothermic reaction. For example, photosynthesis converts carbon dioxide and water to glucose by absorbing sunlight, as
illustrated in Fig. 1.1.8.
Figure 1.1.9 : Energy can transform from one form to another, but energy can neither be created nor destroyed. Source:
https://www.hiclipart.com/free-trans...muhlp/download
The energy conservation law states that energy can transform from one form to another, but the total energy of an isolated
system remains the same.
Energy can transform from one form to another, either through work or heat. An oscillating pendulum is an example of kinetic and
gravitational potential energy periodically transforming into each other through work. Radiation energy in sunlight transfers to
1.1.5 https://chem.libretexts.org/@go/page/371338
chemical energy during photosynthesis, then to heat and work during glucose metabolism, but the total energy remains the same, as
illustrated in Fig. 1.1.9.
1.1: Matter and energy is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
1.1.6 https://chem.libretexts.org/@go/page/371338
1.2: What is chemistry?
Chemistry is the study of matter. More specifically, chemistry studies matter's composition, properties, and transformations.
Properties of the matter are of two types, physical properties, and chemical properties, as illustrated in Fig. 1.3.1.
Figure 1.2.1 : Chemical property relates of elemental composition and its changes. Physical property relates to appearance and its
change. Source: https://www.hiclipart.com/free-trans...mnxsm/download
Examples include color, mass, volume, electrical conductivity, and heat conductivities.
Intensive properties
Physical properties that do not depend on the amount of the matter are called intensive properties, e.g., color, density, and heat
conductivity.
Extensive properties
Physical properties that depend on the amount of the substance, like, mass and volume, are called extensive properties.
Physical process
Any process that changes the matter somehow but does not change the elemental composition is called a physical process.
For example, melting solid to liquid or boiling liquid to the gas state are physical processes.
Mixtures can be separated using physical processes based on the differences in the physical properties of the constituents. Fig. 1.3.2
demonstrates that a magnetic material like iron can be separated from a nonmagnetic material like sulfur using a magnet.
The filtration process can separate a heterogeneous mixture of liquid and solid, like sand in water. Water passes through the filter
paper leaving behind the sand particles on the filter.
The distillation process can separate homogeneous mixtures of solid in liquids or liquids in liquids based on the difference in the
boiling points of the components. For example, a homogeneous salt mixture dissolved in water separates by distilling off the water
at its boiling point, leaving behind the solid salt. Distillation can also separate a mixture of two or more liquids if their boiling
points are different, e.g., a distillation of crude oil separates the components based on their boiling points.
Chromatography is another technique often used to separate mixtures. The mixture, e.g., ink, is adsorbed on a stationary phase,
e.g., on a porous paper, and the separation takes place by flowing liquid phase, e.g., water ascending through the capillaries in the
paper. The component of the ink mixture separate because some components have more ability to stay adsorbed in the solid phase
1.2.1 https://chem.libretexts.org/@go/page/371571
and less ability to solubilize in the liquid phase than the other components. Fig. 1.3.2 Illustrates the physical separation processes
described.
Figure 1.2.1 : Filtration (left), distillation of crud oil (middle) and chromatography of chlorophyll and other pigments of a leaf
extract (right). Source: a) Smokefoot / Public domain, b) Crude_Oil_Distillation-fr.svg: Image originale:Psarianos, Theresa knott ;
image vectorielle:Rogilbertderivative work: Utain () / CC BY-SA (http://creativecommons.org/licenses/by-sa/3.0/), c) No machine-
readable author provided. Flo~commonswiki assumed (based on copyright claims). / CC BY-SA
(https://creativecommons.org/licenses/by-sa/2.5)
For example, methane (CH4) in natural gas is combustible -this is a chemical property. It means methane and oxygen change their
elemental composition to become carbon dioxide and water and release heat after ignition.
Chemical process
A process that changes the elemental composition is called a chemical process or a chemical reaction.
For example, photosynthesis is a chemical process that converts carbon dioxide and water to glucose using energy from sunlight.
Chemical equation
A chemical equation represents a chemical reaction in the form of symbols of elements and compounds involved.
Substances consumed in a chemical reaction are reactants, and the substances formed are products. The reactants are written on
the left side, separated by a plus sign, followed by an arrow, and products are on the right side of the arrow, as illustrated in Fig.
1.3.3. For example, the following chemical equation represents the combustion of methane.
CH4 + 2 O2 ⟶ CO2 + 2 H2 O
Note that the chemical formula without any preceding number, e.g., O2 and CO2, represents one molecule or a unit amount of the
chemical. The number preceding the formula is called the coefficient, and it represents the number of particles or the number of
units involved. For example, the coefficient of 2 in 2H2O in the above chemical equation represents two molecules of water formed
or two moles of water formed, where the mole is the unit amount. Note that the chemical composition has changed in the chemical
reaction. Before, one substance was carbon and hydrogen atoms in a 1:4 ratio, and the other substance was oxygen atoms. After the
reaction, one substance is carbon and oxygen atoms in a 1:2 ratio, and the other is hydrogen and oxygen in a 2:1 proportion.
The physical state of matter is sometimes shown in a chemical equation by the following symbols: (s) for solid, (l) for liquid, (g)
for gas, and (aq) for a substance dissolved in water, as illustrated in Fig. 1.3.3.
1.2.2 https://chem.libretexts.org/@go/page/371571
Figure 1.2.3 : Components of a chemical equation.
The scientific method starts with making observations, giving a tentative explanation, i.e., hypotheses, testing the hypothesis,
i.e., experiment, and deducing a conclusion from the investigation. A truth found through repeated experiments becomes a law,
and a comprehensive explanation of related findings gathered over time becomes a theory.
These scientific method steps are illustrated in Fig. 1.3.4 and described below.
Observation
Observation is the active acquisition of information from a primary source. For example, you fill the air in a car tire and notice that
the pressure reading on the gauge increases as more gas fills the tire. This is an example of observation.
Hypothesis
A hypothesis is a tentative explanation of the observation or a law based on available scientific knowledge. For example, John
visits a friend and starts sneezing. The friend says I have a cat, and you might be allergic to cats. This tentative explanation of
John's sneezing is a hypothesis.
Experiment
Experiments test the hypothesis. For example, John visits another friend who has a cat to figure out whether he is allergic to cats or
not. If he sneezes in this experiment, it supports the hypothesis. If he does not sneeze, the experiment disproves the hypothesis.
1.2.3 https://chem.libretexts.org/@go/page/371571
Conclusion
A hypothesis proven true becomes a conclusion. The hypothesis is rejected or revised if the experiment results do not support it.
For example, scientists from all over the world and in different periods attempted to convert other metals to gold and failed every
time. It concluded that elements do not transform into the more simple matter by any physical or chemical reaction.
Law
If an observation is universally true in repeated experiments, it becomes a law. Examples of law are the following.
1. The pressure of any gas is directly proportional to the amount of gas if temperature and volume are kept constant, is Avogadro's
law.
2. The proportion of atoms of different elements in a compound is always the same, is a law of constant proportion.
3. Mass before any chemical reaction is the same as after the chemical reaction, i.e., mass is conserved in any chemical reaction or
process, which is a law of conservation of mass.
Theory
It is called a theory if someone proposes a comprehensive explanation based on scientific principles to explain several laws and
conclusions on a related topic. For example, the knowledge gathered over time on the properties of matter led Dalton to put
forward Dalton’s atomic theory.
These postulates explain the properties of the matter described in the previous sections. For example, elements can not convert to
simpler substances by any physical or chemical process because they are composed of one type of atom. Atoms are indivisible
according to the first postulate. Compounds can convert to elements by the chemical reaction because the atoms in the compounds
can separate and recombine according to the fourth postulate.
1.2.4 https://chem.libretexts.org/@go/page/371571
Law of multiple proportions
When elements form compounds, the proportions of the elements in those chemical compounds can be expressed in small
whole-number ratios.
For example, hydrogen and oxygen atoms can mix in a 2:1 ratio to provide water (H2O). Still, they can also combine in a 2:2 ratio
to give hydrogen peroxide (H2O2), a different compound. Similarly, carbon and hydrogen combine in a 1:4 ratio to make methane
(CH4); they can combine in a 2:6 ratio to make another ethane compound (C2H6).
1.2: What is chemistry? is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
1.2.5 https://chem.libretexts.org/@go/page/371571
1.3: Measurements
Measurements are an essential part of making observations needed to develop science. Several measurements are commonly done
in everyday life, as illustrated in Fig. 1.3.1. The measured values have two components: a number and a unit.
Figure 1.3.1 : Measurements of mass, length, time, and temperature are part of daily life. Source: https://www.hiclipart.com/free-
trans...xnlrr/download
Numbers
The numbers are composed of digits. The digits are 0, 1, 2, 3, 4, 5, 6, 7, 8, and 9. The digits are written in a row in the number, e.g.,
123, which means one hundred and twenty-three. The numbers include a decimal point. If the decimal point is not marked, it is
assumed to be present at the right of the number. For example, 123 is 123. by conversion, with the decimal point shown in red font.
Sign of a number
Numbers have signs, either +ve or -ve, to the left of a number, e.g., -23.4 and +430. By conversion, no sign means +ve. The signs
are relative to zero; the -ve sign means the number is less than zero, and the +ve sign means the number is more than zero.
1.3.1 https://chem.libretexts.org/@go/page/372550
The number and the sign in calculations
The rules for the sign in a calculated answer are the following.
1. When two positive numbers add, the answer has a +ve sign, e.g., 3+2 = 5.
2. When two negative numbers add, the answer has –ve sign, e.g., -4 + (-2) = -6.
3. When two numbers having opposite signs add, subtract the smaller number from the larger number, and the answer has the
sign of the larger number. For example, -5 +3 = -2.
4. In subtraction, change the sign of the subtracted number and then follow the addition rules. For example, subtract 3 from 5:
5-(+3) = 5-3 = 2. Note that 3 is subtracted, and its sign changed before operating addition. Another example: subtract -6
from 2: 2-(-6) = 2+6 = 8
5. When two positive numbers multiply, the answer has a +ve sign, e.g., 2x3 = 6.
6. When two negative numbers multiply, the answer has a +ve sign, e.g., (-4) x (-3) = 12.
7. When the two numbers multiplied have opposite signs, the answer has a –ve sign, e.g., (-3) x 2 = -6 and 4 x (-4) = -16.
8. When a number is divided by another number, it follows multiplication rules for the sign. For example,
= −3 .
−4 4 −4 9
= 2, = 2, = −2, and
−2 2 2 −3
Percentage calculations
The percentage (%) is the part out of a hundred, as illustrated in Fig. 1.3.3. The percentage is calculated as part divided by the total
and then multiplied by a hundred, i.e.:
Part amount
percentage % = × 100
Total amount
Example 1.3.1
Example 1.3.2
A piece of 18K green color gold jewelry has 7.5 g gold, 2.0 g silver, and 0.5 g copper. Calculate the percentage of gold in the
jewelry?
Solution
Part = 7.5 g gold, Total = 7.5 g gold + 2.0 g silver + 0.5 g copper = 10 g jewelry.
Plug in values in the formula and calculate:
Part amount 7.5ggold
percentage = × 100 = × 100 = 75%gold
Total amount 10 g
1.3.2 https://chem.libretexts.org/@go/page/372550
Writing numbers in scientific notation
Sometimes the given number is too large or too small to be easily written, read, and grasped. Scientific notation is one approach to
changing a too large or a too-small number into an easily readable and writable number. The following steps convert a given
number to scientific notation.
1. Move the decimal point to the right or the left side, one digit at a time, till the largest non-zero digit becomes one's place.
For example, move the decimal in 12,700,000 seven times to the right to obtain 1.27, and move the decimal in
0.000,006 six-time to the left to get 6. The numbers 1.27 and 6 obtained are the coefficients of the scientific
notation.
2. The coefficient is multiplied by 10x, where x is a power of ten. The power of ten equals the number of times the decimal
moved. The sign of the power is +ve if the decimal moved to the left and -ve if the decimal moved to the right. For
example, 12,700,000 in scientific notation is 1.27 x 107, and 0.000,006 is 6 x 10-6.
Units
Physical properties like mass, length, and temperature are measured. The measured value is a combination of a number and a unit,
as illustrated in Fig. 1.3.4. For example, a person's height is 1.83 meters, where 1.83 is a number and a meter is a unit.
Units are quantities defined by the standard that peoples agree to use as a reference.
For example, the meter is defined as the distance light travels in a vacuum in 1
299,792,458
of a second.
Systems of units
There are different sets of units used in different systems of units. For example, 1.83 meters and 6.00 feet show the same length
value but using a unit of ‘meter’ from the international system of units (SI) and ‘foot’ from the English system of units. The
international system of units (SI) is universally used in scientific work. There are seven base units in SI, as listed in Table 1.
Table 1: Base units in the International System of Units (SI).
Measurement Unit Abbreviation
Time second s
Length meter m
Mass kilogram kg
Temperature kelvin K
1.3.3 https://chem.libretexts.org/@go/page/372550
Measurement Unit Abbreviation
Note
The following section is based on the 2019 redefinition of the SI base units: https://en.Wikipedia.org/wiki/2019_redefinition
_of_the_SI_base_units#Kilogram, accessed on May 2nd, 2020
Time
Time is the progress of existence and events that occur in succession from the past through the present to the future.
In old times, the time measuring device was the hour sandglass shown in Fig. 1.3.5. The basic unit of time is second (s), a standard
unit of time in all the measurements systems. Other units of time are minute (min) which is equal to 60 s, and hour which is equal
to 60 min or 3600 s.
Definition: Second
The duration of 9,192,631,770 periods of the radiation corresponding to the transition between the two hyperfine levels of the
ground state of the cesium 133 atoms at a temperature of 0 K.
Figure 1.3.5 : German; Half-hour sandglass; Horology. Source: Metropolitan Museum of Art / CC0
Length
Length is a measure of distance, i.e., a numerical measurement of how far apart the objects or points are.
Fig. 1.3.6. illustrated the concept of length. The SI unit of length is a meter (m).
Meter
299,792,458
of a second.
1.3.4 https://chem.libretexts.org/@go/page/372550
Figure 1.3.6 : Measure of length. Source: https://www.hiclipart.com/free-trans...nhlxx/download
Mass
The mass of an object is a measure of its inertia.
Inertia is the resistance of any physical object to any change in its velocity. Mass determines the strength of the gravitational
attraction of an object to another object -a property commonly used in modern balances for mass measurements, as shown in Fig.
1.3.7. SI unit of mass is the kilogram (kg).
Kilogram
Earlier definition: The mass of one cubic decimeter of water at the melting point of ice.
Current definition: Kilogram (kg) is defined by taking the fixed numerical value of the Planck constant h to be
6.62607015×10−34 when expressed in the unit kg⋅m2⋅s−1.
Temperature
Temperature is a physical property of matter that expresses hotness or coldness, as illustrated in Fig. 1.3.8.
Temperature is a manifestation of the thermal energy of the matter, which is a source of the flow of energy in the form of heat
from a hot object to a cold object when they are in contact with each other.
1.3.5 https://chem.libretexts.org/@go/page/372550
The SI unit of temperature is Kelvin (K).
Kelvin
Kelvin (K) is defined by taking the fixed numerical value of the Boltzmann constant k to be 1.380649×10−23 when expressed
in the unit kg⋅m2⋅s−2⋅K−1.
A 0 K, also called absolute zero, is the temperature of a matter at which no energy can be removed as heat from the matter. The
freezing point of water is 273.15 K, and the boiling point of water is 373.15 K.
Amount of substance
In chemistry, the amount of substance (n) measures the number of specified elementary entities. The elementary particles in
chemistry are usually atoms in the case of elements and molecules or formula units in the case of compounds. SI unit of the amount
of a substance is a mole (mol).
Mole (mol)
Fig. 1.3.9 illustrates one mole of aluminum, copper, and carbon. The molar mass is the mass in grams of one mole of that
substance, i.e., the mass of 6.02214076×1023 atoms or molecules. Usually, the number of particles is shown with four significant
figures, i.e., 6.022 x 1023 atoms or molecules in one mole of the substance.
Figure 1.3.9 : Amount in one mole of aluminum, copper, and carbon shown relative to US Quarter coin.
Source: Public information, photo: R. Press/NIST; graphic design: N. Hanacek/NIST
1.3.6 https://chem.libretexts.org/@go/page/372550
Electric current
Electric current is the flow rate of electric charge past a point or a region.
It could be the flow of electrons in electric wires or the flow of cations and anions in opposite directions as in electrolytes, as
illustrated in Fig. 1.3.10. The SI unit of current is ampere (A).
Figure 1.3.10 : Illustration of the definition and direction of electric current, i.e, the current (I) is the rate of flow charges (Q) per
unit time (t). Source: And1mu/ CC-BY-SA-4.0, https://commons.wikimedia.org/wiki/F...ricCurrent.gif
Luminous intensity
Luminous intensity measures the wavelength-weighted power emitted by a light source in a particular direction per unit solid
angle.
The solid angle is measured in steradian (sr), analogous to the radian. The radian is a planar angel that gives the length of the
circumference of a circle, and the steradian is a 3D angle, like a cone, that gives an area on the surface of a sphere, as shown in Fig.
1.3.11.
Figure 1.3.11 : A solid angle is a three-dimensional analog of a circular angle that relates a portion of the volume of a sphere to the
surface area it subtends. If that area equals the sphere’s radius squared, the solid angle is one steradian. This diagram displays two
solid angles of one steradian, viewed from different directions. Source: Andy Anderson / CC BY-SA
(https://creativecommons.org/licenses/by-sa/4.0)
SI unit of luminous intensity in a given direction is Candela (cd).
Candela (cd)
Candela (cd) is defined by taking the fixed numerical value of the luminous efficacy of monochromatic radiation of frequency
540×1012 Hz, Kcd, to be 683 cd⋅sr⋅W−1, or cd⋅sr⋅kg−1⋅m−2⋅s3, where W is watt –a SI unit of power described by kg.m2.s-3.
Prefixes in SI
In several situations, the measured number with the base unit is either too large or too small. For example, a person's height is
comfortable to represent in the meter as the height is usually in a 1 m to 2 m range. However, the diameter of the earth, i.e.,
12,700,000 m, and the diameter of red blood cells, i.e., 0.000,006 m, are too large and too small, respectively. The unit needs to be
revised so that the number with it is easy to read and write.
Prefixes are used in SI to increase or decrease the base unit by order of tens.
1.3.7 https://chem.libretexts.org/@go/page/372550
For example, kilo (k) means a thousand times, i.e., 1 km means 1000 m and 1 kg means 1000 g. Similarly, micro (µ) means one-
millionth time, i.e., 1 µm is 10-6 m, and 1 µg is 10-6 g. Table 2 lists commonly used prefixes in SI.
Table 2: Commonly used prefixes in SI (note: m means “meter” in SI units, but as a prefix, it means “mili”)
Prefix Means Abbreviation
Gega 109 G
Mega 106 M
Kilo 103 k
Deci 10-1 d
Centi 10-2 c
Mili 10-3 m
Micro 10-6 µ
Nano 10-9 n
Pico 10-12 p
A new unit may be defined and used if there is no appropriate prefix available in SI for some specific type of measurement. For
example, the diameter of atoms varies in the range of 1 x 10-10 m to 5 x 10-10 m, where the prefix pico (p, 10-12) is too small, and
the prefix nano (n, 10-9) is large. A new unit called angstrom (Å) is defined as 1Å = 10-10 m for reporting atomic diameter and
inter-atomic distances.
Derived units
The units in SI other than the seven base units are Derived units obtained by combining the base units.
For example, The SI unit of volume is meter-cube (m3), equal to the space occupied by a cube of 1m on each edge, as illustrated in
Fig. 1.3.12.
Figure 1.3.12 : Relative volume in m3, dm3 and cm3. Diagram of a 10x10x10 cubes. The small cubes represent one decimeter-cube
(dm3); the big cube represents one meter-cube (1 m3). The purpose is to show that even though there is 10 dm in 1 m, there are
1000 dm3 in 1 m3. More generally, it shows that the conversion factor between units of volume is the cube of the conversion factor
between corresponding length units. Source: Download for free at https://openstax.org/details/books/chemistry
Usually, the volume is reported in decimeter-cube (dm3), commonly known as liter (L). One liter is a volume occupied by a cube
that is one dm on each edge. Another commonly used unit of volume is the centimeter-cube (cm3), which is also called cc or mL.
One mL is the volume occupied by a cube that is one cm on each edge. The dm3 is a thousandth of m3, and cm3 is the thousandth
of dm3, i.e., 1000 dm3 = 1 m3, and 1000 cm3 = 1 dm3, as illustrated in Fig. 1.3.11.
1.3.8 https://chem.libretexts.org/@go/page/372550
Relationship of SI units with metric and English system of units
SI was developed from the metric system. Some basic units are different, but both systems have much in common, using the same
prefixes. The English system of units uses a different set of units except for the common unit of time. Table 3 compares the
standard measuring units in the three systems of measurement.
Table 3: Common measurement units in three conventional systems of measurements
Quantity English unit Metric unit SI unit Relationships
1 kg = 2.205 lb
Mass Pound (lb) Gram (g) Kilogram (kg)
1 kg = 1000 g
0.946 L = 1 qt
Volume Quart (qt) Liter (L) Cubic meter (m3)
1 m3 = 1000 L
°F = (1.8 × °C) + 32
Temperature Degree Fahrenheit (°F) Degree Celsius (°C) Kelvin (K)
K = °C + 273.15
1.3: Measurements is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
1.3.9 https://chem.libretexts.org/@go/page/372550
1.4: Significant Figures
Significant figures are related to errors associated with the measured numbers. It is important to understand the significant figures
because when calculations are made using numbers with errors, the answer cannot have less error than the error in any original
number. The answer needs correction for the significant figures.
Figure 1.4.1 : Representation of high accuracy with less precision (left) and low accuracy with high precision (right). Source:
DarkEvil / Public domain
Systematic errors
Systematic errors are constant, i.e., they have the same value in every measurement. For example, meter rod is a little short or a
little long than a meter, it will introduce a systematic error. Systematic errors usually happen due to inaccurate calibration of the
measuring instrument. The systematic errors determine how much the measured value differs from the actual value.
Random errors
Random errors are the statistical variability of the measured number. Random errors vary from one observation to another. Random
errors cancel out if many measurements are taken and averaged. Scientific measurements are usually taken at least in triplicate and
averaged to minimize random errors. The random errors determine how close the repeat measured numbers are to each other.
Accuracy
Accuracy or trueness of the measurement is defined as how close the average value is to the actual value.
The closer the average is to the actual value, the more accurate or true it is, as illustrated in Fig. 1.4.1. The trueness depends on
systematic errors, i.e., less systematic error, more accurate the average.
Precision
Precision is defined as how close the individual measurements are to each other.
The closer the individual values are to each other, the more precise the measurement is, irrespective of whether it is accurate or not,
as illustrated in Fig. 1.4.1. Precision depends on random errors, i.e., more substantial random errors mean less precision.
That is, there is no error in it. For example, a purchase of one dozen oranges contains exactly 12 nos of oranges; it can not be 11.5
or 12.5.
1.4.1 https://chem.libretexts.org/@go/page/372551
For example, when the same one dozen oranges are purchased by mass, the balance may read it 1572.6 g, or 1573 g, or 1570 g,
depending on whether the smallest digit that the balance displays is 0.1g, 1 g, or 10 g. Suppose the balance is accurate to 1 g and
reports the mass 1573 g; the actual mass may be anywhere in the range of 1572.5g-to-1573.4g. The smallest measured digit, i.e.,
the number in one's place, in this case, is an estimated number associated with an error. By convention, the estimated digit has ±1
errors associated with it. For example, the above-mentioned measured numbers are reported in science as 1572.6 g ± 0.1 g, 1573 g
± 1 g, or 1570 g ± 10 g, respectively. The estimated digits are shown in bold fonts in the examples.
The smallest digit in the display of digital instruments is an estimated number. In measurement using instruments that do not have a
digital display, the smallest digit marked on the instrument plus one digit less than the minimum marked digit is added to the
reported value. The smallest reported digit is an estimated digit. For example, the length of the pencil in Fig 1.4.2 is reported as
17.7 cm using the ruler on the bottom, where 17 includes the smallest digit marked on the ruler, and the last digit, i.e., 0.7 is an
estimated digit. By convention, the error range in this value is shown as 17.7 ± 0.1. The same length is 17.70 cm using the ruler on
the top in Fig. 1.4.2l, where 17.7 includes the smallest digit marked on the ruler, and the last reported digit, i.e., 0, is an estimated
digit. By convention, the error range in this value is shown as 17.70 ± 0.01. The estimated digits are marked in bold fonts.
Figure 1.4.1 : Illustration of measurements that yield inexact numbers. The length of a blue rectangle is 7.60 ± 0.01 cm by using a
more accurate ruler on the top and 7.6 ±0.1 cm by using a less accurate ruler on the bottom. The estimated digit is marked in bold
fonts.
Significant figures
All the digits reported in the measured value, including the estimated digit, are significant figures (SF).
For example, 1572.6 g, 1573 g, and 1570 g have significant figures of 5, 4, and 3, respectively.
Caution
Note that zero in the last reading 1570g is not significant; it is a placeholder zero that is needed to place the estimated digit 7 at
tens place.
It is crucial to find significant figures in measured numbers because, when they are used in calculations, the answer cannot have
less error than the maximum error in any measured number used in the calculation. The rules to determine the significant numbers
in a measured number are the following.
1. All non-zero digits are significant, e.g., 1572 has 4 SFs. The zeros may or may not be significant. In the following
examples, the zeros in bold fonts are nonsignificant.
2. Zeros between non-zero digits are significant, e.g., 1305.6 has 5 SFs.
3. Leading zeros are not significant, e.g., 0.0134 has 3 SFs.
4. Trailing zeros are not significant if there is no decimal point present, e.g., 1570 has 3 SFs. Trailing zeros are significant if
the decimal point is present, e.g., 1570. has 4 SFs because the decimal point is present. Similarly, 0.0124 has 3 SFs, but
0.01240 has 4 SFs because the decimal point is present.
5. Confusion arises when more than one trailing zeros and the decimal point is absent. For example, 1500 g has 2 SFs by
convention, but if the balance was accurate to 10 g, one of the zero was an estimated digit and was significant. Converting
the number to a scientific notation resolves this issue. The coefficient part of the scientific notation shows all the significant
figures in the measurement. For example, the number 1500 g, if shown in scientific notation as 1.5 x 103 has 2 SFs, but the
same number shown as 1.50 x 103 has 3 SFs.
1.4.2 https://chem.libretexts.org/@go/page/372551
Rounding the calculated answer involving inexact numbers
When inexact numbers are used in calculations, the answer needs to be rounded to an appropriate number of significant figures,
determined by the following rules.
Rules of rounding
1. A number is rounded by keeping the larger digits equal to significant figures and dropping or replacing the remaining
smaller digits with placeholder zeros. The placeholder zeros are in bold fonts in the following examples. For example,
13543 becomes 13500 when rounded to three significant.
2. If the largest digit dropped is 4 or less than 4, it is simply dropped. For example, 23145 becomes 23100, when rounded to
three significant figures.
3. If the largest digit dropped is 5 or more than 5, then the smallest digit retained is increased by one. For example, 13543
becomes 14000 when rounded into two significant figures.
In the following rules, the track of significant figures that dictate the significant figures in the answer is kept by using bold
fonts.
1. In addition and subtraction, the answer has the same number of decimal places as the number with the smallest number of
decimal places in the original numbers. For example, 13.2 + 12.252 = 25.452 is rounded to 25.5 to keep one decimal place.
2. In multiplication and division, the answer has the same number of significant figures as the original number with the
smallest number of significant figures. For example, 1.35 x 2.1 = 2.835 is rounded to 2.8.
Note
If mathematical operations are performed in a series of steps, keep track of the significant figures but do not round off
intermediate answers. Carry as many digits as possible from the intermediate answers to the next calculation step. Round off
the final answer following the above rules. For example, (13.2 + 12.252) x (1.35 x 2.1) = 25.452 x 2.835 = 72.15642 is round
to 72 in agreement with 2.835 that should have been rounded to 2.8 in a one-step calculation. Rounding the intermediate
answers will lead to incorrect final answer of 71 instead of more correct 72, i.e. (13.2 + 12.252) x (1.35 x 2.1) = 25.5 x 2.8 = 71
Caution
Exact numbers have an unlimited number of significant figures, which means they do not restrict the significant figures in the
calculated answer.
Example 1.4.1
1572.6 g/12 = 131.05 g
Explanation: The answer has 5 SFs because 12 is a counted number and exact. The only inexact number in the calculation that
dictates the significant figures in the solution is 1572.6, which has 5 SFs.
Example 1.4.2
One dozen oranges were were sold 11 times. Calculate the total oranges sold?
Solution
12 × 11 = 132 oranges
1.4.3 https://chem.libretexts.org/@go/page/372551
Explanation: The answer is not rounded because both the numbers in the calculation are exact, so the answer is also exact with
unlimited significant figures.
1.4: Significant Figures is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
1.4.4 https://chem.libretexts.org/@go/page/372551
1.5: Unit conversions
Conversion of values in one unit to the same value in another unit, as illustrated in Figure 1.5.1 is often needed in scientific
calculations. The unit conversion includes the following.
Conversion of the same type of measurement in the same system of measurement, e.g., conversion of a measured value of the
length in meters to kilometers in SI;
conversion of the same type of measurement in different systems of measurements, e.g., conversion of a measured value of the
length in kilometers from SI to miles in the English system; and
conversion of one type of measurement to another type of measurement, e.g., conversion of a measured value of the mass in g
to volume in mL of a substance using the density of the substance.
Conversion factors
The conversion factors are derived from equality between the given unit and the desired unit. For example, 1 cm = 10-2 m is
equality between centimeter and meter. The conversion factors are derived from the equality by the following steps.
−2
Both sides of the equality are divided by one side to get one conversion factor. For example, 1 =
10
1 cm
m
, which is a conversion
factor for cm to m.
Then both sides are divided by the other side of the equality to get the second conversion factor. For example, 1 = 1 cm
−2
, which is
10 m
10
−2
m
1 cm
1.83 m × = 183 cm
−2
10 m
Keep track of the units that cancel out and the unit left in the answer. If all the units cancel out, leaving only the desired unit means
the chosen conversion factor is correct. An incorrect conversion factor leads to an unwanted unit in the answer, e.g., in the above
calculation if incorrect conversion factor is chose, it will lead to:
−2 2
10 m m
1.83 m × = 0.0183
1 cm cm
, where no unit is canceled and the answer has units that are not the desired. It means an incorrect conversion factor was employed.
1.5.1 https://chem.libretexts.org/@go/page/372552
−2
1 cm 10 m
and
−2
10 m 1 cm
Some of the common qualities in SI are listed in Table 1 and the English system in Table 2.
Note
The prefixes are exact numbers. The equalities within the same system of measurement are exact numbers. Therefore the
equalities and the conversion factors derived from them are exact numbers. Significant figures in the answers involving exact
and inexact numbers are dictated by inexact numbers only.
Example 1.5.1
Note that in this example, the answer has the significant figures the same as in the given number because the conversion factor
within the same system of measurement is numbers.
1 km = 1,000 m 1 kg = 1,000 g 1 L = 10 dL
1 cm = 10 mm 10-1 1 dL = 100 mL
1 mL = 1 cm3 = 1 cc
ft = 12 in. 1 lb = 16 oz 1 qt = 4 cups
1 mi = 5,280 ft 1 qt = 32 fl oz
1 gal = 4 qt
These equalities between different systems usually have one side in the equality, which is the number 1, as exact, while the
other side is considered an inexact number.
For example, 1 kg = 2.205 pounds (lb) has an exact number (1 kg) on the left side but an inexact number (2.205 lb) with 4 SFs on
the right. Remember that only the inexact numbers dictate the significant figures in the answer.
There are some exceptions to the above general rule. Some equalities between units in different systems are defined and
considered exact. They are stated to be exact in the reference tables.
For example, 1 inch = 2.54 cm is defined, which means both sides are the exact number.
Table 3: some of the common SI-to-English system equalities
1.5.2 https://chem.libretexts.org/@go/page/372552
Length Mass Volume
factor used to convert the volume to the mass of a substance. Reciprocal density, i.e., = , is the second conversion factor that 1
d
v
Example 1.5.2
The density of ethanol at 20 oC is 0.7893 g/mL; what is the mass of 10.0 mL of ethanol?
Solution
Multiply the given volume with the conversion factor that has volume in the denumerator, i.e., \(frac{m}{v} to get the mass
desired.
0.7893 g
10.0 mL × = 7.89 g
1 mL
Note that 1mL is exact, and 10.0 and 0.7893 are inexact numbers with 3SF and 4 SF, respectively. The answer has 3 SFs.
Example 1.5.3
The density of gold is 19.30 g/mL; what is the volume of 10.123 g of gold?
Solution
Multiply the given mass with the conversion factor that has a mass in the denumerator, i.e., f racvm to get the mass desired.
1 mL
10.123 g × = 0.5245 mL (1.5.1)
19.30 g
For calculations in chemistry, the number of moles of a substance is considered equal to its coefficient in a balanced chemical
equation.
For example:
2H +O ⟶ 2H O
2 2 2
2 2
1 mol O 2 mol H O
2. equality: 1 mole O2 = 2 mole H2O, conversion factors: 2 mol H O
2
and 1 mol O
2
and
2 2
2 mol H 2 mol H O
3. equality: 2 moles H2 = 2 moles H2O, conversion factors: 2 mol H O
2
and 2 mol H
2
.
2 2
The use of these conversion factors in the calculation is explained in the following example.
1.5.3 https://chem.libretexts.org/@go/page/372552
Example 1.5.4
If 5 moles of oxygen (O2) is consumed, how many moles of water are produced by the chemical equation mentioned above?
Solution
Given: 5 mol O2, Desired: ? mol H2O
Multiply the given quantity with the conversion factor that has the given unit in denumerator and the desired unit in the
numerator:
2 mole H2 O
5.0 mole O2 × = 10 mole H2 O
1 mole O2
Example 1.5.5
1 g µ
1.00 × 10
−2
g × = 1.00 × 10 g
4
µ
−6
10 g
Note that the first conversion factor converts mg to g, and then the second conversion factor converts g to the desired unit µg.
The same calculation can be done using two conversion factors in a rwo:
−3
10 g µ
1 g
10.0 mg × × = 1.00 × 10 g
4
µ
−6
10 mg 10 g
Example 1.5.6
Note that the first conversion factor converts km to m, and then two conversion factors are needed to convert h to s via min.
Example 1.5.7
A prescription says a dosage of 0.225 mg of Synthroid to be taken once a day. If tablets in stock contain 75 µg of Synthroid,
how many tablets are needed per day?
Solution
1.5.4 https://chem.libretexts.org/@go/page/372552
The given mass is in mg, while the equality "1 tablet = 75 µg" takes mass in µg. First convert mg to base unit, i.e., g, then from
g to needed unit, i.e., µg, and finally take appropriate conversion factor from the two given by the equality to convert µg to
tablet, i.e., three conversion factors in a row:
10
−3
g 1 gµ 1 tablet
0.225 mg × × × = 3.0 tablet
−6
1 mg 10 g 75 µg
Example 1.5.8
A healthy person has 16% body fat by mass. Calculate the mass of fat in kg of a person who weighs 180. lb?
Solution
Given: mass of a person = 180. lb, Desired: mass of body fat in kg.
16% body fat by mass means: 16 lb body fat = 100 lb body mass, and the equality kg and lb is: 1 kg = 2.20 lb. Take one
conversion factor from each equality such that the units cancel out leaving the desired unit in the answer:
16 lb body fat
1 kg body fat
180. lb body mass × × = 16 kg body fat
100 lb body mass 2.20 lb body fat
Note that there are three inexact numbers in the calculation, i.e., 180, 16, and 2.20, and the answer as two significant figures in
agreement with the smallest significant figure among the inexact numbers.
1.5: Unit conversions is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
1.5.5 https://chem.libretexts.org/@go/page/372552
1.6: Equations and graphs
Equalities where both sides have a single term, i.e., monomial, lead to the conversion factors. If one or both sides of the equality
have more than one term, i.e., polynomial, it leads to a formula that does the same job, i.e., to convert units. Temperature
conversion equations and ideal gas equations and the manipulation procedure of these equations are described below as examples
of how to manipulate the questions needed for making calculations.
Figure 1.6.1 : Relations among temperature scales in common use. Source: MikeRun / CC BY-SA
(https://creativecommons.org/licenses/by-sa/4.0)
Celsius (oC)
The Celsius scale has O oC at the freezing point of water and 100 oC at the boiling point of water. Celsius is the base unit of
temperature in the metric system.
TK = TC + 273,
where TK is the temperature in Kelvin, and TC is the temperature in degrees Celsius. This equation converts temperature in Kelvin
to temperature in Celsius.
A 0 K, also called absolute zero, is the temperature of a matter at which no energy can be removed as heat from the matter. There is
no negative temperature on the Kelvin scale.
Fahrenheit (oF)
Fahrenheit is the base unit of the English system, with 32 oF at the freezing point of water and 212 oF at the boiling point of water.
Fahrenheit is times shorter and shifted up by 32 than Celsius. So the relationship between the two is:
5
9
TF = × TC + 32,
5
where TF is the temperature in Fahrenheit, and TC is the temperature in degrees Celsius. This equation converts temperature in
Celsius to temperature in Fahrenheit.
1.6.1 https://chem.libretexts.org/@go/page/372553
Addition or subtraction of the same number on the two sides of an equation does not change the equality. Subtracting 32 from both
sides of the above equation leads to:
9
TF − 32 = × TC +32 −32 ,
5
9
TF − 32 = × TC .
5
Multiplication or division by the same number on both sides of an equation does not change equality. Remember that multiplication
or division should apply to every term on either side of the equality. Enclose the side with more than one term in small brackets and
then do the multiplication of division operation so that it applies to each term in the bracket. Multiplying both sides of the above
equation with leads to:
5
5 5 9
× (TF − 32) = × × TC
9 9 5
5
× (TF − 32) = TC
9
Swapping the sides of an equation does not change equality. Swapping the sides in the above equation to bring TC to the left:
5
TC = × (TF − 32)
9
TK = TC + 273,
TK − 273 = TC ,
and finally swap the left and right side to bring TC to the left:
TC = TK − 273.
P V = nRT ,
where P is pressure, V is volume, n is the amount of gas in moles, T is the temperature (in K), and R is the proportionality constant
called an ideal gas constant. Dividing both sides of the equation with V leads to:
V nRT
P × = ,
V V
nRT
P = .
V
It allows calculating the pressure of a gas sample if the amount in moles, temperature in kelvin, and volume of the gas sample are
known, along with the value of the constant R in the consistent units. Similarly, rearranging the equation leads to formulas for
calculating, V, T or n of a gas sample:
nRT PV PV
V = , T = , and n =
P nR RT
1.6.2 https://chem.libretexts.org/@go/page/372553
Graphs
The graph is a visual presentation of a relationship between two variables. Fig. 1.6.2 shows a graph that presents a relationship
between the volume and pressure of a given amount of gas at a constant temperature, known as Boyle’s law.
Figure 1.6.2 : Boyle’s law –the relationship between pressure and volume of a given amount of gas at a constant temperature.
The typical components of a graph are the following.
1. A title that tells what the graph is about, e.g., “Boyle’s Law: ” in the graph of Fig. 1.6.2.
2. Axes: the x-axis is a horizontal line, and the y-axis is a vertical line. Axes usually have an evenly distributed scale starting
from zero. The x-axis represents the independent variable, and the y-axis represents the dependent variable. For example,
volume is independent, and pressure is the dependent variable in Fig. 1.6.2.
3. Axes labels that tell the variable's name and the units of measurement. For example, volume (in3) and pressure (in Hg),
where in3 and in Hg are the units of the variables.
4. Symbols representing experimental points. For example, ∆ symbols in Fig 1.6.2, at the crossing of a vertical line starting
from an experimental value independent variable on the x-axis and a horizontal line starting from the corresponding value
of the dependent variable on the y-axis.
5. A curve connects the experimental points and shows the trend in the relationship. For example, the cure in Fig. 1.6.2 tells
that the pressure decreases as the volume increases.
Interpretation of a graph
Interpretation is reading the trend or relationship between the variables plotted. For example, Fig. 1.6.2 shows that the pressure
decreases as the volume of gas increases. The curve also allows reading the value of one variable from the value of the other. For
example, if the volume is 30 in3, the pressure would be ~50 in Hg. To read: draw a vertical line from the given value on the x-axis
and a horizontal line from the point where the vertical line crosses the curve. Then, read the value where the horizontal line meets
the y-axis. The process is reversed when the given value is of the dependent variable on the y-axis, and the desired is the
corresponding value of the independent variable on the x-axis. For example, if the pressure is 40 in Hg, the volume is about 35 in3.
Fig. 1.6.3 is another example of a graph that shows a relationship between the molar mass and density of the gases at a constant
temperature. The cure in this graph tells that the density of gases increases as the molar mass increases.
Figure 1.6.3 : Curve showing gases having lower molar masses are less dense at the same temperature.
1.6: Equations and graphs is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
1.6.3 https://chem.libretexts.org/@go/page/372553
1.7: Density and specific gravity measurements
Density
Density is the mass-to-volume ratio of a substance.
Density is a physical characteristic of matter. Each substance has a characteristic density that can be used as one hint in identifying
a substance.
Gases have very low density, usually expressed in g/L. For example, air density is around 1.224 g/L at sea level and 15 oC. The
density of liquids and solids is usually expressed in g/mL. For example, the density of water at 4 oC is 1.00 g/mL.
Objects that are less dense than water float, and the denser objects than water sink in the water. For example, oil is less dense than
water and floats on water. Metals are denser than water and sink in water. The density of some common substances is listed in
Table 1
Table 1: Densities of some common substances
Substance Density (g/mL)
hydrogen 0.000089
water 1.00
magnesium 1.74
aluminum 2.70
iron 7.86
copper 8.92
silver 10.50
lead 11.34
mercury 13.59
gold 19.30
Density measurement
Density (d) is calculated from the mass (m) and volume (V) of a substance by the formula:
m
d =
V
Mass is usually measured using an analytical balance. The volume of liquids can be measured using a graduated cylinder, pipet, or
density bottle. The volume of regular solids can be calculated from the geometric parameters. For example, the volume of a
rectangle is equal to length x width x height. The volume of a cube is equal to the edge length cubed.
The volume of an irregular shaped sold is usually measured because the substances that are denser than water sink and displace an
equal amount of water. Fig. 1.7.1 illustrates the density measurement of an irregular-shaped solid object that sinks in water, as
explained in the following example.
1.7.1 https://chem.libretexts.org/@go/page/372554
Figure 1.7.1 : Measuring mass, volume, and density of an irregular solid object. Source: modified images from
https://www.hiclipart.com
Example 1.7.1
Caution
Most of the time, Calculators give more significant numbers and sometimes less than needed; both need correction. In example
1.7.1, the calculator displays 8, i.e., one significant figure, but three zeros are added to make four significant figures.
Figure 1.7.2 : Illustration of Osteoporosis as a result of bone density decrease. Source: BruceBlaus / CC BY-SA
(https://creativecommons.org/licenses/by-sa/4.0)
Specific gravity
Specific gravity is the ratio of the object's density to the density of water, i.e.:
Density of an object
Specific gravity =
Density of water
1.7.2 https://chem.libretexts.org/@go/page/372554
Specific gravity is the ratio of the object's density to the density of water, i.e.:
Density of an object
Specific gravity =
Density of water
The units cancel out in the ratio. Therefore, the specific gravity is a unitless number. The density of water is 1.0 g/mL at room
temperature, so the specific gravity is equal to the density of the object expressed without a unit.
When substances dissolve in water, the density of the solution is usually different from pure water. For example, the density of
whole blood for humans is ~1.060 g/mL. The density of urine varies in the range of 1.0050 g/mL to 1.030 g/mL. Both the blood
and urine have dissolved substances in water that increase the density from that of pure water. Both high and low density or
specific gravity than the normal range of urine indicates medical problems. An increase in the specific gravity of urine indicates
that it is due to an increase in the solutes caused by dehydration, diarrhea, or infection. Similarly, a decrease in solute concentration
decreases the specific gravity of urine, which indicates medical problems like renal failure.
Figure 1.7.3 : Schematic drawing of an API hydrometer combined with a thermometer. Source: Milton Beychok, CCO
1.7: Density and specific gravity measurements is shared under a Public Domain license and was authored, remixed, and/or curated by
Muhammad Arif Malik.
1.7.3 https://chem.libretexts.org/@go/page/372554
1.8: Heat and its measurements
Heat
Heat is energy in transfer to and from a thermodynamic system by a mechanism other than the work or the transfer of matter.
Energy exists in different forms, but energy transforms from one form to another through work or heat. In chemical reactions, a
significant portion of energy transfer happens through heat. Green leaves in plants trap the energy from light and store it as
chemical energy in the form of glucose synthesized from water and carbon dioxide, as illustrated in Fig. 1.8.1.
Figure 1.8.1 : Photo-synthesis traps the light energy and stores it as chemical energy. Source: https://www.hiclipart.com/free-
trans...gzitr/download
A reverse process happens when we eat food, i.e., the food substances are converted to carbon dioxide, water, and energy, as
illustrated in Fig. 1.8.2. Some energy is used to maintain our body temperature at an average level. Another portion is used to drive
chemical reactions that consume energy and synthesize substances we need.
Figure 1.8.2 : Basic overview of energy and human life. Source: Mikael Häggström / Public domain
Body Temperature
The human body continually loses energy as heat to the environment. The heat released by exothermic reactions in the food
digestion process supplies the heat needed to maintain the body temperature at an average level. Human body temperature varies
over a small range, as illustrated in Fig. 1.8.3. Hypothermia is <35.0 °C (95.0 °F), normal body temperate is 36.5–37.5 °C (97.7–
99.5 °F), fever and hyperthermia is >37.5 or 38.3 °C (99.5 or 100.9 °F), Hyperpyrexia is >40.0 or 41.0 °C (104.0 or 105.8 °F).
1.8.1 https://chem.libretexts.org/@go/page/372555
Figure 1.8.3 : Variation in human body temperate. Source: Foxtrot620 / CC BY-SA (https://creativecommons.org/licenses/by-
sa/4.0)
Measurement of heat
Temperature is a manifestation of the thermal energy of an object, but the temperature and the energy are not the same.
Temperature is a measure of hotness or coldness, and it is intensive, and energy is an extensive physical property of matter. For
example, 1 g of water has some energy, but 2 g of water at the same temperature has twice the energy.
Specific heat
The calory (c or cal) is a non-SI unit of heat, work, and energy.
The heat needed to raise the temperature of 2 g of water by 1 oC is 2 calory, and to raise the temperature of 2 g of water by 2 oC is 4
calory. That is, heat energy (q) is directly proportional to both the mass (m) and change in temperature (∆T) of an object:
q ∝ mΔT
Introducing a constant of proportionality, i.e., specific heat (Cs) changes the proportionality to equality, that is known as the heat
equation:
q = Cs mΔT
The heat energy needed to raise the temperature of 1g of a substance by 1 oC (or 1K) is specific heat if the substance.
1 cal = 4.184 J (exact)
1.8.2 https://chem.libretexts.org/@go/page/372555
The energy unit in nutrition and food is food calory or Calory (C) that is written with capital C, and it is equal to 1000 cal, i.e.,
g ⋅o C
)
Al 0.902
C (graphite) 0.720
Fe 0.451
Cu 0.385
Au 0.128
H2O(steam) 2.00
Wood 1.76
Concrete 0.88
Glass 0.84
heat or the heat energy needed for a given temperature change of a material. A coffee cup calorimeter illustrated in Fig. 1.8.4 is
commonly used. It consists of a coffee cup with a lid. It contains water in which the hot or cold object is immersed, or a chemical
reaction is performed for heat exchange with the water. A wire loop is used to stir the water. A thermometer records the change in
the temperature of the water. The heat absorbed or released by the water is calculated from the data of mass, specific heat, and
temperature change of the water.
1.8.3 https://chem.libretexts.org/@go/page/372555
Figure 1.8.4 : Photo of coffee cup calorimeter set. Source: Community College Consortium for Bioscience Credentials / CC BY
(https://creativecommons.org/licenses/by/3.0)
Example 1.8.1
Immersing a hot object in 50.0 grams of water in a coffee cup calorimeter increased the water temperature from 22.0 oC to 28.8
o
C. How much heat is gained or lost by water?
Solution
Given: m = 50.0 g water, Cs = 4.184 (\frac{J}{g \cdot^{o} C}\), ∆T = Tf – Ti = 28.8 oC – 22.0 oC = 6.8 oC, Desired: heat
gained or lost by water, i.e, qwater = ?
Calculations:
J ∘
qwater = Cs mΔT = 4.184 × 50.0 g × 6.8 C = +1400 J
∘
g C
Example 1.8.2
Ammonium nitrate (NH4NO3) is used as a cold pack in hospitals to decrease the temperature of a targeted skin area. When
3.21 g of NH4NO3 is dissolved in 50.0 g of water in a coffee cup calorimeter at 25.0 oC, the temperature decreases to 20.4 oC.
What is the amount of heat absorbed or released by water?
Solution
Given: m of solution = 50.0g H2O + 3.21 g NH3NO3 = 53.2 g, Cs = 4.184 (\frac{J}{g \cdot^{o} C}\), ∆T = Tf – Ti =20.4 oC -
25.0 oC = -4.6 oC, Desired: heat gained or lost by the solution, i.e, qsolution = ?
Calculations:
J
∘ 3
qsolution = Cs mΔT = 4.184 × 53.2 g × (−4.6 C) = −1.0 × 10 J
∘
g C
1.8.4 https://chem.libretexts.org/@go/page/372555
Note
In the problems involving the heating of solution, the mass of solution includes the mass of water and the solute. It is assumed
that the specific heat and density of the dilute solution are the same as that of pure water.
Example 1.8.3
Calcium chloride (CaCl2) is used as a hot pack, i.e., to raise the temperature of a targeted portion of the skin. When CaCl2
dissolves in water, heat is released by the process and absorbed by the surrounding water, increasing the temperature of the
water.
When 5.00 g of CaCl2 is dissolved in 50.0 g of water at 23oC in a coffee cup calorimeter, the temperature rises to 39.2oC. What
is the amount of heat absorbed or released by water?
Solution
Given: m of solution = 50.0g H2O + 5.00 g CaCl2 = 55.0 g, Cs = 4.184 (\frac{J}{g \cdot^{o} C}\), ∆T = Tf – Ti =39.2 oC -
23.0 oC = 16.2 oC, Desired: heat gained or lost by the solution, i.e, qsolution = ?
Calculations:
J ∘ 3
qsolution = Cs mΔT = 4.184 × 55.0 g × (16.2 C) = 3.7 × 10 J
∘
g. C
, where ∆Ho is the enthalpy of reaction, i.e., the heat evolved or absorbed at constant pressure conditions. This reaction releases
2803 kJ of energy as heat per mole of glucose. The –ve sign of ∆Ho tells that the reaction releases the energy, i.e., the reaction is
exothermic. The same amount of energy is released when glucose is digested and converted to the same products in the living
system under the same conditions.
1.8.5 https://chem.libretexts.org/@go/page/372555
Figure 1.8.1 : Bomb calorimeter. Source: Lisdavid89, CC BY-SA 3.0 <https://creativecommons.org/licenses/by-sa/3.0>, via
Wikimedia Commons
Fuel value
The heat of combustion per g of a food is called the fuel value of the food
Glucose is one of the carbohydrates. Carbohydrates are food components that are compounds of carbon, hydrogen, and oxygen in
which the ratio of hydrogen-to-oxygen is 2-to-1. The average fuel value of carbohydrates is 17 kJ/g.
The second major class of food is fat. For example, the chemical equation of tristearin C57H110O6, which is a fat, is:
∘
2C H O (s) + 16 O (g) ⟶ 114 CO (g) + 110 H O(l) ΔH = −71609 kJ
57 110 6 2 2 2
The –ve sign with energy value tells the energy is released, or the reaction is exothermic. The average fuel value of fats is 38 kJ/g.
The third primary type of food is protein. Protein is mainly used as a building material in living systems, but it is also used as a
source of energy. The average fuel value of protein is 17 kJ/g.
The average fuel values of the three major food components are listed in Table 2. These values are used to calculate the fule value
of food servings, as explained in the following examples.
Table 2: Average fuel value of food components
Food Fuel value kJ/g
Carbohydrates 17
Proteins 17
Fats 38
Example 1.8.1
Calculate the energy released from one cup of orange juice that contains 26 g of carbohydrates, no fat, and 2 g of protein.
Solution
Carbohydrate: 26 g × 17 kJ/g = 442 kJ
Protein: 2 g × 17 kJ/g = 34 kJ
Total energy released = 476 kJ
The energy value in kJ can be converted to food calories (C) by using the conversion factor based on the following relations.
1.8: Heat and its measurements is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
1.8.6 https://chem.libretexts.org/@go/page/372555
1.9: Heat and changes in physical states of matter
Among the four physical states of matter, solid has the lowest thermal energy. Intermolecular forces in solids are strong and do not
let the molecules slide past each other. The molecules and the bonds in them can still have vibrational motions that account for the
thermal energy contents of the material.
The temperature reflects the thermal energy content of the material—the addition of heat increase the vibrational motions, and
temperature increases. Ultimately, the solid changes to a liquid and the liquid changes to a gas phase as more heat is added, as
illustrated in Figure 1.9.1.
Figure 1.9.1 : Illustration of the relationship between energy and phase changes of matter. Source: https://www.hiclipart.com/free-
trans...poitg/download
The heat of fusion is usually expressed in the units of joules per gram ( ) for the unit amount in grams or in joules per mole (
J
g
J
mol
)
for the unit amount in moles.
If heat is removed from a substance at its melting point, the reverse of melting, i.e., freezing, happens, i.e., the liquid gradually
changes from liquid to solid phase. The energy equal to the heat of fusion is released during the freezing process. Fig. 1.9.2 shows
ice and water at 0 oC –an example of melting and freezing.
1.9.1 https://chem.libretexts.org/@go/page/372556
Figure 1.9.2 : Ice and water at 0 oC –an example of melting and freezing. Source: Ulflund / CC0
The energy needed to evaporate a unit amount of a liquid is called the heat of vaporization (∆Hvap).
The heat of vaporization is usually expressed in the units of joules per gram ( ) for the unit amount in grams or joules per mole (
J
mol
) for the unit amount in moles.
The reverse of evaporation is called condensation, which releases heat equal to the heat of vaporization. Fig. 1.9.3 demonstrates the
co-existence of liquid and gas-phase bromine at room temperate through the simultaneous evaporation and condensation processes.
When the temperature reaches the boiling point of the liquid, the temperature does not increase further, but the added heat is used
to evaporate the liquid. Heating increases the temperature of the gas phase after all of the liquid has changed to the gas phase.
Figure 1.9.3 : Bromine at room temperature –an example of evaporation and condensation. Source: Alchemist-hp (pse-
mendelejew.de) / CC BY-SA 3.0 DE (https://creativecommons.org/licenses...3.0/de/deed.en)
The energy required in sublimation (∆Hsub) is the addition of the heat of fusion and the heat of vaporization, i.e.,:
The reverse of the sublimation is called deposition, i.e., the gas phase changes directly to the solid phase. Fig. 1.8.4 shows the
sublimation of iodine crystals on a hot plate and deposition of iodine gas on an ice-cold watch glass.
1.9.2 https://chem.libretexts.org/@go/page/372556
Figure 1.9.4 : Iodine crystals on hot plate sublime by the heat and the gaseous iodine deposits on ice-cold watch galss. Source:
Alvy16 / CC BY (https://creativecommons.org/licenses/by/4.0)
The sublimation is responsible for drying clothes below 0 oC conditions in cold areas. Sublimation is also used in freeze-drying
vegetables and other foods. Bacteria can not grow on dried foods because they need some moisture to grow. Fig. 1.9.5 shows the
terminologies related to the phase changes described in the previous paragraphs.
Heating curve
A graphical presentation of the relationship of heat added versus the temperature change and phase changes of a matter is
called a heating curve.
Fig. 1.9.6 shows the heating curve of water. The curve shows the heating of ice initially, followed by co-existing of solid and liquid
at the freeing point, then hating of liquid water, then co-existing of liquid and gas phases at the boiling point, and finally the heating
of steam –the gas phase of water. The reverse of the heating curve is called the cooling curve.
1.9.3 https://chem.libretexts.org/@go/page/372556
Figure 1.9.6 : Heating curve of water. Source: Community College Consortium for Bioscience Credentials / CC BY
(https://creativecommons.org/licenses/by/3.0)
Example 1.9.1
Calculate the energy required to heat 10.0 g of ice from -20.0 oC to steam (water vapor) at 110 oC?
Solution
1st step –heating the ice from -20.0 oC to the melting point of the ice, i.e., 0.00 oC:
m = 10.0 g, Cs of ice = 2.06 J
∘
g⋅ C
, ∆T = 0.00 oC – (-20.0 oC) = 20.0 oC
J ∘
q1 = Cs mΔT = 2.06 × 10.0 g × 20.0 C = 412 J
∘
g⋅ C
2nd step – melting of ice, multiply the heat of fusion with the amout of substance:
m = 10.0 g, ∆Hfus = 334 J
g
.
J
q2 = ΔHfus × m = 334 × 10.0 g = 3340 J
g
3rd step –hating of the water from 0.00 oC to the boiling point of water, i.e., 100 oC:
m = 10.0 g, Cs of liquid water = 4.184 J
∘
g⋅ C
, ∆T = 100 oC – 0.00 oC = 100 oC
J
∘
q3 = Cs mΔT = 4.184 × 10.0 g × 100 C = 4180 J
∘
g⋅ C
4th step – boiling of liquid water, multiply the heat of vaporization with the amount of the substance:
m = 10.0 g, ∆Hvap = 2260 J
J
q4 = ΔHvap × m = 2260 × 10.0 g = 22600 J
g
g⋅ ∘ C
, ∆T = 110 oC – 100 oC = 10.0 oC
1.9.4 https://chem.libretexts.org/@go/page/372556
J ∘
q5 = Cs mΔT = 2.00 × 10.0 g × 10.0 C = 200 J
∘
g⋅ C
1 kJ
30700 J × = 30.7 kJ
1000 J
Note
The most significant portion of the heat is consumed in boiling the water to steam, i.e., 22.6 kJ out of 30.7 kJ total. The same
amount of heat is released when the steam condenses. That is why the steam burn is much more severe than the burn by hot
water.
1.9: Heat and changes in physical states of matter is shared under a Public Domain license and was authored, remixed, and/or curated by
Muhammad Arif Malik.
1.9.5 https://chem.libretexts.org/@go/page/372556
CHAPTER OVERVIEW
2: Elements
2.1: Dalton’s atomic theory
2.2: Subatomic particles and a modern view of an atom
2.3: Atoms of elements
2.4: The periodic table
2.5: Electrons in atoms
2.6: The periodic trends in properties of the elements
2: Elements is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
1
2.1: Dalton’s atomic theory
Atom
Atom is the smallest particle of an element that retains the element's characteristics.
For example, gold is an element. Magnifying a section of the gold surface large enough would look like a packing of atoms, as
illustrated in Fig. 2.1.1.
Figure 2.1.1 : A piece of gold (right) is an element that is composed of particles called atoms, as shown in the model (left). Source:
https://www.hiclipart.com/free-trans...ghixq/download.
Dalton's atomic theory was the first significant attempt to explain the basic knowledge of atoms gained over time. Postulates of
Dalton's atomic theory are the following.
Dalton's atomic theory is the basis of the current atomic theory, though the atoms are no more considered ‘indivisible.’ According
to the current knowledge, subatomic particles like electrons, protons, and neutrons compose the atoms. However, the subatomic
particles do not represent the element.
2.1: Dalton’s atomic theory is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
2.1.1 https://chem.libretexts.org/@go/page/372863
2.2: Subatomic particles and a modern view of an atom
Atoms are composed of fundamental subatomic particles: electrons, protons, and neutrons.
Electron was the first subatomic particle discovered. The discovery of electrons is related to the study of cathode rays and the basic
knowledge of charges, i.e., there are two types of charges +ve and –ve; like charges repel each other; opposite charges attract each
other as illustrated in Fig. 2.2.1; electric and magnetic field deflects the moving charges.
Figure 2.2.1 : This diagram describes two equal (like) point charges repelling each other and two opposite charges attracting each
other, with an electrostatic force F which is directly proportional to the product of the magnitudes of each charge and inversely
proportional to the square of the distance between the charges. Ke is Coulomb's constant. Source: File:CoulombsLaw.svg:
User:Dna-Dennis / *derivative work RJB1 / CC BY (https://creativecommons.org/licenses/by/3.0)
Figure 2.2.2 : Cathode ray tube with cathode-ray shown deflected by an electric field. Source: Kurzon / Public domain
2.2.1 https://chem.libretexts.org/@go/page/372864
Figure 2.2.3 : Plum pudding model of an atom. Atoms were initially thought to contain many hundreds or thousands of electrons as
shown in this schematic representation of the plum pudding model. Source: Tjlafave / CC BY-SA
(https://creativecommons.org/licenses/by-sa/4.0)
Figure 2.2.4 : Illustration of Rutherford’s gold foil experiment (left). Interpretation of the results – Almost all the atom's mass is in a
tiny nucleus space. α -particles that come close to the atom’s nucleus are deflected (right). Source: https://www.hiclipart.com/free-
trans...xivja/download and https://www.hiclipart.com/free-trans...txlqr/download
2.2.2 https://chem.libretexts.org/@go/page/372864
The discovery of the neutron
The mass of protons and electrons did not account for an atom's overall mass, which led to a search for another subatomic particle.
James Chadwick discovered that bombarding α -rays on a beryllium target produced a highly penetrating radiation consisting of a
beam of neutral particles, now called neutrons. The presence of neutrons in the nucleus accounts for the missing mass of atoms.
Other sub-atomic particles have been discovered, e.g., quarks that constitute protons and neutrons, but their knowledge is not
critical for understanding basic chemistry.
Figure 2.2.5 : depiction of the atomic structure of the helium atom. The darkness of the electron cloud corresponds to the line-of-
sight integral over the probability function of the 1s atomic orbital of the electron. The magnified nucleus is schematic, showing
protons in pink and neutrons in purple. Source: User:Yzmo / CC BY-SA (http://creativecommons.org/licenses/by-sa/3.0/)
Atomic mass unit (amu) which is 1/12 of the mass of a single carbon atom that has 6 protons and 6 neutrons in it.
−27
1 amu = 1.660539606660(50) × 10 kg
Angstrom(Å)
1 Å = 10-10 m
These units are usually used for masses, charges, and diameters of atoms. Table 1 lists the basic properties of the subatomic
particles.
Table 1: Basic properties of subatomic particles
Particle Charge (e) Mass (amu)
Proton +1 1.0073
Neutron 0 1.0078
2.2.3 https://chem.libretexts.org/@go/page/372864
Particle Charge (e) Mass (amu)
2.2: Subatomic particles and a modern view of an atom is shared under a Public Domain license and was authored, remixed, and/or curated by
Muhammad Arif Malik.
2.2.4 https://chem.libretexts.org/@go/page/372864
2.3: Atoms of elements
Symbol of an element
Elements are represented by symbols, the first alphabet of their English or non-English name, written in capital letters. For
example, C for carbon, N for nitrogen, and I for iodine. Usually, another alphabet is also chosen from the element's name and
written as a small letter, e.g., Al for aluminum, Zn for zinc, and Ca for calcium. Some element symbols are derived from non-
English names, e.g., Na for sodium from the Latin name natium, Cu for copper from Latin cupurum, and Ag for silver from Latin
argentum.
Atoms of the same element have the same number of protons, and atoms of different elements have different numbers of protons.
In other words, the number of protons in an atom defines the element. There are 118 elements known at this time; the number of
protons in atoms varies from 1 for hydrogen to 118 for Oganesson (Og).
Atomic number
The atomic number defines the element. A subscript to the left of the symbol of an element represents the atomic number. For
example, H shows one proton in a hydrogen atom, and C shows 6 protons in a carbon atom.
1 6
Mass number
The number of protons plus the number of neutrons in an atom is the mass number.
A superscript to the left of the symbol of an element represents the mass number. For example, H is a hydrogen atom with atomic
1
1
number 1, mass number 1, and no neutrons, while F is a fluorine atom with 9 protons and 10 neutrons.
19
9
Number of electrons
The number of electrons in an atom equals the number of protons minus the charge on the atom.
The number of electrons is equal to the number of protons in the case of a neutral atom, as there is no charge on a neutral atom.
Cation
A neutral atom can lose some electrons and become a positively charged particle, called a cation.
The charge is represented as a superscript on the right side of the element symbol, e.g. H is hydrogen without any electron, i.e.,
1
1
+
1 proton, 0 neutrons, and 0 electrons. Ca is calcium with two fewer electrons than protons, i.e., 20 protons, 20 neutrons, and
20
40
2+
18 electrons.
Anion
An atom can gain electrons and become a negatively charged particle, called an anion.
For example, O is oxygen with two more electrons than protons on it, i.e., 8 protons, 8 neutrons, and 10 electrons. A F is
16
8
2− 19
9
−
fluorine with one more electron than protons on it, i.e., 9 protons, 10 neutrons, and 10 electrons. Fig. 2.3.1 illustrates the gain or
loss of electrons from neutral atoms.
2.3.1 https://chem.libretexts.org/@go/page/372865
Figure 2.3.1 : Comparison of subatomic particles in neutral atoms and ions formed from them after the gain or loss of an electron.
Source: modified from: Wdcf / CC BY-SA (https://creativecommons.org/licenses/by-sa/3.0)
and I is an integer equal to charge number and charge the sign of the charge number: + or -.The number of protons, neutrons, and
electrons is calculated by using the following formulas:
Number of protons = A,
Number of neutrons = A – Z, and
Number of electrons = Z – (charge I),
where a charge is a + or - sign of the charge number.
Example 2.3.1
8
O
1+
: number of protons = Z = 8, number of neutrons = A – Z = 16-8 = 8,
and number of electrons = Z – (charge I) = 8 – (+1) = 8 – 1 = 9.
Note
If charge number I is 1 in X , it is usually not written, but a number more than one is written. For example,
A
Z
I charge 16
8
O
−
has
change = -1, O has charge charge -2, and O has charge = +1.
16
8
2− 16
8
+
Isotopes
All atoms of the same element have the same number of protons but can have a different number of neutrons. For example, 1
1
H ,
H , and H have neutrons equal to 0, 1, and 2, respectively.
2 3
1 1
Atoms of the same element that have a different number of neutrons are called isotopes.
For examples, H , H , and H are isotopes of hydrogen illustrated in Fig. 2.3.2. Another example is Li, and Li are two isotopes
1
1
2
1
3
1
6
3
7
3
of lithium. Natural samples of elements usually have almost constant ratios of isotopes. Table 1 lists some Isotopes of elements and
their percent abundance in typical natural samples.
2.3.2 https://chem.libretexts.org/@go/page/372865
Figure 2.3.2 : Isotopes of hydrogen. Source: Dirk Hünniger; Derivative work in english - Balajijagadesh / CC BY-SA
(https://creativecommons.org/licenses/by-sa/3.0)
Table 1: Some of the important isotopes of elements with abundance in typical natural samples
Element Isotopes Abundance (%)
1
1H 99.99
Hydrogen 2
1H 0.01
3
1
H Negligible
6
Li 7.6
Lithium 3
7
3
Li 92.4
12
6C 98.93
Carbon 13
6C 1.07
14
6C Negligible
35
17 Cl 75.78
Chlorine 37
17 Cl 24.22
79
Br 50.69
Bromine 35
81
35
Br 49.31
235
U 0.72
Uranium 92
238
92
U 99.28
Atomic mass
The atomic mass listed in the periodic table is the weighted average of the masses of the isotopes present in a natural sample
of the element.
, where ∑ means summation over all isotopes of the element, the fractional abundance of the isotope is the % abundance divided
by 100. Fig. 2.3.3. illustrates how the atomic mass is listed in a periodic table.
Figure 2.3.3 : Two isotopes are present: copper-63 (62.9 amu) and copper-65 (64.9 amu), in abundances 69% + 31%. The standard
atomic weight for copper is the average, weighted by their natural abundance. Source: DePiep / CC BY-SA
(https://creativecommons.org/licenses/by-sa/4.0)
2.3.3 https://chem.libretexts.org/@go/page/372865
Example 2.3.2
Calculate the atomic mass of chlorine with two isotopes in nature samples, i.e., 35
17
Cl of mass 34.969 amu and % abundance
75.78% and Cl of mass 36.996 amu and % abundance 24.22%.
37
17
Solution
Formula: Atomic mass = ∑[( mass of isotope ) × ( fractional abuncance of the isotope )
Plug in the given values in the formula and calculate:
75.78 24.22
(34.969 amu × ) + (36.996 amu × ) = 35.45 amu
100 100
Note
1. Atomic masses of isotopes are close to but not the same as their mass numbers. For example, Cl has mass number mass =
35
17
35, but the atomic mass of this isotope is 34.969 amu as shown in the above example.
2. The weighted average atomic mass is usually closer to the mass number of the most abundant isotope, e.g., 35.45 amu in
the above example is close to the mass number 35 of Cl isotope, which is the most abundant isotope.
35
17
3. The periodic table reports the atomic mass as calculated in the above example, i.e., the weighted average of the masses of
the isotopes present in the natural sample of the element.
2.3: Atoms of elements is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
2.3.4 https://chem.libretexts.org/@go/page/372865
2.4: The periodic table
Early developments
The discoveries of elements happened over a long time. As the list of known elements grew, scientists tried to arrange them based
on their properties. Mendeleev arranged elements in a table based on atomic masses. It resulted in elements having similar
properties placed next to each other in most cases. There were gaps intentionally left in the table for the elements that were
predicted based on the knowledge from the periodic table but not yet discovered.
Few exceptions were there observed where the properties of the elements did not agree with the group in which they were placed
based on their atomic masses. Mosely developed a method to measure atomic numbers based on X-ray spectroscopy. The
arrangement of the elements based on the atomic number instead of atomic masses removed the discrepancies in Mendeleev’s
periodic table.
Figure 2.4.1 : Periodic table of elements. Fort color of element symbols: blue are gases, red are liquids and black are solids at room
temperature. Source: Modified from https://www.exceltemplates.org/category/academic)
Periods
The periodic table has seven horizontal rows called periods. The periods are numbered: 1 at the top to 7 at the bottom.
1. The 1st period has only two elements: hydrogen in group 1 and helium in group 18, with a gap from group 2 to group 17.
2.4.1 https://chem.libretexts.org/@go/page/372866
2. The 2nd and the 3rd periods have eight elements each, filling groups 1 and 2 followed by groups 13 to 18, leaving a gap
from group 3 to group 12.
3. The periods 4th and 5th periods have eighteen elements that are filled successively from group 1 to group 18.
4. The periods 6th and 7th have 32 elements, each: the first two in groups 1 and 2, the next fourteen elements in separate rows
below the table. These two rows of 14 elements each are called Actinides and Lanthanoids, respectively. Then the next
sixteen elements fill the groups 3 to 18.
Groups
The periodic table has 18 vertical columns called groups or families. The groups are numbed starting from 1 on the leftmost and
going through to 18 at the rightmost.
Alkali metals
The 1st group is called alkali metals. The alkali metals include lithium (Li), sodium (Na), potassium (K), rubidium (Rb),
cesium (Cs), and francium (Fr), shown in Fig. 2.4.2. The alkali metals are the most reactive among the metals in the periodic
table. They react vigorously with water, as shown in Fig. 2.4.3.
Figure 2.4.2: Alkali metals, from left to right: lithium, sodium, potassium, rubidium, and cesium. Source: Tomihahndorf at
German Wikipedia, Dnn87 Contact email: [email protected], and http://images-of-elements.com/potassium.php.
Figure 2.4.3 : Potassium reacting with water. Source: Ozone aurora / Philip Evans / CC BY-SA (https://creativecommons.org
/licenses/by-sa/3.0)
Hydrogen
Hydrogen is in group 1 but is not included in alkaline earth metals. Hydrogen is a nonmetal and has properties quite different
from alkali metals or any other group of elements.
The 2nd group is called alkaline earth metals. It includes beryllium (Be), magnesium (Mg), calcium (Ca), strontium (Sr),
barium (Ba), and radium (Ra). They are reactive metals but less reactive than alkali metals. Alkaline earth metals impart
characteristic color to a flame. Salts of alkali metals are used in firework formulation to give distinctive colors to the firework,
as shown in Fig. 2.4.4.
2.4.2 https://chem.libretexts.org/@go/page/372866
Figure 2.4.4 : New years eve fire works at Al Majaz waterfront, Sharjah UAE. Metal salts including alkali metals are used to give
distinctive colors in fireworks. Source: Fariz Safarulla / CC BY-SA (https://creativecommons.org/licenses/by-sa/3.0)
Transition metals
Groups 3 to 12 are called transition metals. They include precious metals like gold, silver, platinum, and construction metals
like iron. Some make catalysts and are found in enzymes and other bio-molecules, like hemoglobin and chlorophyll.
Group 13 to 16
Group 13 to group 16 does not have a unique name. They comprise nonmetals at the top and metals at the bottom of each
group called post-transition metals. Important nonmetals include carbon, nitrogen, oxygen, phosphorous, and sulfur.
Halogens
Group 17 elements are called halogens. The halogens include fluorine (F), chlorine (Cl), bromine, iodine (I), and astatine (At).
The halogens are highly reactive nonmetals. Chlorine is gas, bromine is liquid, and iodine is solid at room temperature, as
shown in Fig. 2.4.5.
Figure 2.4.5 : From left to right: chlorine, bromine, and iodine. Source: W. Oelen / CC BY-SA
(https://creativecommons.org/licenses/by-sa/3.0)
Noble gases
Group 18 is called noble gases. They include helium (He), neon (Ne), argon (Ar), krypton (Kr), xenon (Xe), and radon (Rn).
They are the least reactive of all the elements. Noble gases are used to create an inert atmosphere for chemical reactions. Noble
gases are also used in the lighting system because of their chemically inert nature , as illustrated in Fig. 2.4.6.
2.4.3 https://chem.libretexts.org/@go/page/372866
Figure 2.4.6 : Glowing noble gases in glass vials with low pressure inside. Power supply: 5 kV, 20 mA, 25 kHz. Source: New work
Alchemist-hp (talk) www.pse-mendelejew.de); original single images: Jurii, http://images-of-elements.com. / CC BY
(https://creativecommons.org/licenses/by/3.0)
Metalloids
The dividing line between metals and nonmetals is a staircase line starting from B and ending at At. The elements on the
5 85
staircase line are metalloids except for aluminum and polonium, which are considered metals. Metalloids have properties in-
between metals and nonmetals; e.g., they have moderate heat and electrical conductivity.
Nonmetals
Elements towards the top-right corner of the periodic table and hydrogen are called nonmetals.
1. The nonmetals usually have properties opposite to metals, e.g., they are not typically shiny, brittle if solid, and poor
conductors of heat and electricity.
2. Nonmetals tend to make ionic compounds by accepting electrons from metals and making molecular compounds by
reacting with each other.
Two groups in nonmetals also have unique names, i.e., group 17 is called halogens, and group 18 is called Noble gases.
Generally, alkali metals are the most reactive, followed by alkaline earth metals, and halogens are the most reactive nonmetals.
Noble gases are the least reactive nonmetals, also called inert gases.
2.4.4 https://chem.libretexts.org/@go/page/372866
Figure 2.4.7 : Gold nugget –a metal, pure silicon –a metalloid, and sulfur –a nonmetal (from left to right). Source: Gold: Rob
Lavinsky, iRocks.com – CC-BY-SA-3.0 / CC BY-SA (https://creativecommons.org/licenses/by-sa/3.0), Silicon: Enricoros at
English Wikipedia / Public domain, Sulfur: Benjah-bmm27 / Public domain
Figure 2.4.8 : A table of elements important to human life. Pink elements are essential that are macroneutrients; blue elements are
important that are microneutrients. Source: Tosaka / Public domain
2.4.5 https://chem.libretexts.org/@go/page/372866
Allotropes
Different forms of the same element in the same physical state are called allotropes.
The allotropes are different structural modifications of the element. For example, carbon exists in several allotropic forms; two of
them are shown in Fig. 2.4.9. Another example is O2, and O3 are gaseous forms of oxygen.
Figure 2.4.9 : Diamond and graphite are two of the allotropes of carbon: pure forms of the same element that differ in structure.
Source: Diamond_and_graphite.jpg: User:Itubderivative work: Materialscientist / CC BY-SA
(http://creativecommons.org/licenses/by-sa/3.0/)
Figure 2.4.10 : Metal ions in a sea of electrons. Source: Steven Legg (Sjlegg) / Public domain
2.4: The periodic table is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
2.4.6 https://chem.libretexts.org/@go/page/372866
2.5: Electrons in atoms
The electronic structure of an atom determines the properties of the element. Knowledge of electromagnetic radiation is described
first as it plays an essential role in understanding the electronic structure of atoms.
Figure 2.5.1 : Illustration of electric field, magnetic field, propagation direction and wavelength of an electromagnetic wave.
Source: DECHAMMAKL / CC BY-SA (https://creativecommons.org/licenses/by-sa/4.0)
Electromagnetic radiations
Electromagnetic radiations are waves that are oscillating electric and magnetic fields. The wave propagates in one direction, e.g.,
along the x-axis. The electric field oscillates perpendicular to it, e.g., along the y-axis. The magnetic field oscillates perpendicular
to both, e.g., along the z-axis, as illustrated in Fig. 2.5.1. The distance between two consecutive maxima or between any two
consecutive same phase points along the wave is called wavelength (λ , pronounced ‘lambda’). The number of waves that pass a
reference point in one second is called frequency (ν , pronounced ‘nu’). The speed of electromagnetic radiation is called the speed
of light (c). The speed of light is the product wavelength and frequency, i.e.,
c = λν
The speed of light (c) is 3.00 x 108 m/s in a vacuum. The energy (E) of electromagnetic radiation is directly proportional to
frequency, i.e.,
E = hν
λ
shows that the energy is inversely proportional to the
wavelength.
hc
E =
λ
Fig. 2.5.2 illustrates the range of electromagnetic radiations that differ from each other concerning wavelength, frequency, or
energy. The separation of the radiations based on their wavelength gives a spectrum. Visible light is a small portion of the
spectrum of electromagnetic radiations, as illustrated in Fig. 2.5.2.
2.5.1 https://chem.libretexts.org/@go/page/372867
Figure 2.5.2 : A spectrum of electromagnetic radiations. The visible range of the spectrum is expanded with the wavelength shown
in nm units. Source: Philip Ronan, Gringer / CC BY-SA (https://creativecommons.org/licenses/by-sa/3.0)
Figure 2.5.3 : Continues spectrum of visible light (400-700 nm range, top) and line spectrum from hydrogen (bottom). Source:
it:Utente:Sassospicco / Public domain and Merikanto, Adrignola / CC0
The shell
Quantum numbers determine the allowed energy values of an electron in an atom.
Principal quantum number (n) can have any integer value starting from 1, i.e., 1, 2, 3, 4, and so on.
The smaller the n, the lower is the energy state, and the closer the electron is to the nucleus, the more tightly held the
electron is by the nucleus.
The value of n defines the shell, i.e., 1st shell has n = 1, 2nd shell has n = 2, 3rd shell has n =3, and so on.
Bohr introduced this concept of quantization of electronic energy levels. Fig. 2.5.4 illustrates Bohr’s atomic model.
2.5.2 https://chem.libretexts.org/@go/page/372867
Figure 2.5.4 : Bohr model of the atom generalized (left), and electron energy levels of a hydrogen atom with energy values and
spectral emission series. Source: Brighterorange / CC BY-SA (http://creativecommons.org/licenses/by-sa/3.0/), and
https://www.hiclipart.com/free-trans...cnxhb/download
When an electron jumps from a lower shell to a higher shell, it absorbs electromagnetic radiation of energy equal to the
energy gap between the initial and the final shell.
When an electron jumps from a higher shell to a lower shell, it emits radiation equal to the energy gap between the initial
and the final shell.
Fig. 2.5.3 illustrates the emission of radiation from atoms –it is a line spectrum because only discrete energy levels, called shells,
are allowed to electrons in an atom.
Figure 2.5.5 : The shapes of the first five atomic orbitals are: (from left to right) 1s, 2s, 2px, 2py, 2pz, and of the five togather with
the nucleus at the center. Source: Neon_orbitals.JPG, public domain, and John Trombley / CC BY
(https://creativecommons.org/licenses/by/4.0)
The subshell
A second quantum number, called Azimuthal quantum number (l) defines subshells within a shell.
The subshells are usually designated as s, p, d, f, ...
Each shell has subshells equal to the shell number. For example, 1st shall have only one subshell, i.e., s. It is designated 1s,
where the number is the principal quantum number, and the letter is the subshell. The 2nd shell has two subshells 2s and 2p; the
3rd shell has three subshells 3s, 3p, and 3d; and the 4th shell has four subshells 4s, 4p, 4d, and 4f.
The energy order of subshells is 1s<2s<2p<3s<3p<4s and so on.
Fig. 2.5.6 helps in remembering the order. This figure is drawn by placing the orbitals in columns and shell numbers in rows in
increasing order of n from top to bottom, starting from 1s orbitals in the first column and first row, p orbitals in the second, d in the
third, and f in the fourth. The filling of electrons follows arrows going from corner to corner, starting from the top left corner of the
topmost cell, as described in more detail in a later section.
2.5.3 https://chem.libretexts.org/@go/page/372867
Figure 2.5.6 : Order of filling of electrons in subshells. Start from the top and follow the arrows.
Orbital
The orbital is the region in space around the nucleus of an atom where electrons are most likely found.
Each subshell has a certain number of orbitals in them. The orbitals have a specific shape and orientation. The s subshell has only
one orbital spherically symmetrical, like a ball with a nucleus at the center. The 1s orbital is smaller than 2s, and 2s is smaller than
3s, but they all have a spherically symmetrical shape, as illustrated in Fig. 2.5.5. The p subshell has three orbitals. Each p orbital is
a dumbbell shape with two lobes, i.e., px oriented along the x-axis, py along the y-axis, and pz along the z-axis, as illustrated in Fig.
2.5.5.
Degenerate orbitals
All the three p orbitals in the same shell have the same energy, i,e, 2px, 2py, and 2pz is a set of degenerate orbitals. The d subshell
has five degenerate orbitals, and f subshell has seven degenerate orbitals, as shown in Fig. 2.5.7. Their shapes and orientations are
more complex and not shown here.
2.5.4 https://chem.libretexts.org/@go/page/372867
Figure 2.5.7 : Relative energies of shells, subshells, and orbitals in multi-electron atoms. Source: https://www.hiclipart.com/free-
trans...yfbwx/download
The energy order of orbitals is 1s<2s<2p<3s<3p<4s and so on, as shown in Fig. 2.5.6 and Fig. 2.5.7.
2.5.5 https://chem.libretexts.org/@go/page/372867
have the electron configuration 1s22s22p3, 1s22s22p4, 1s22s22p5, and 1s22s22p6, respectively. Atomic number 3 to 10, i.e., lithium to
neon, completes the 2nd-row.
The outermost shell is called the valence shell, and electrons in the valence shell are called valence electrons. 1st shell is the
valence shell for 1st-row elements hydrogen and helium. 2nd-row elements have an inner shell with configuration 1s2, and a
valence shell containing 2s and 2p being filled. The inner shell is also called the core-shell, and the electrons in the core-shell
are called the core electrons.
Figure 2.5.8 : The electron configuration of the first twenty elements in the periodic table of elements. Valence electrons are shown
in red color fonts.
2.5: Electrons in atoms is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
2.5.6 https://chem.libretexts.org/@go/page/372867
2.6: The periodic trends in properties of the elements
Properties of elements generally show a periodic trend that correlates with their position in the periodic table. The properties and
their periodic trends are described below.
Valence electrons
Electrons in the outermost shell are the valence electrons. Fig. 2.5.8 shows the valence electrons of the first twenty elements in red
fonts.
All elements in a group have the same number of valence electrons equal to the first digit of their group number. For example,
1st group of hydrogen and alkali metals has one valence electron, 2nd group of alkali metals has two valence electrons,
halogens in 17th group have seven valence electrons, and noble gases in 18th group have eight valence electrons.
Caution
The transition metals in groups 3 to 12, and inner-transition metals, i.e., lanthanoids and actinoids -the two rows of elements
placed below the periodic table, are the exception to the general trend of valence electrons described above. Valence electron
configurations of transition metals and inner-transition metals are not described here; it is beyond the scope of this book.
The valence electrons mainly determine the chemical properties of the elements. The elements in the same group have similar
chemical properties because they have the same valence shell electron configuration. The elements in a row show a gradual change
in chemical properties because their valence shell electron configuration changes gradually along the row.
Lewis symbols
1. Lewis symbols show the valence electrons as dots around the symbol of an element. One dot represents one valence
electron, e.g., .
2. The dots are shown on any of the four sides of the symbol.
3. A single dot on the top, bottom, left, or right is shown four valence electrons. Then start pairing the dots beyond four
valence electrons, as shown in Fig. 2.6.1 for the first twenty elements.
4. Helium is an exception that has only two valence electrons, but they are shown paired.
Figure 2.6.1 : Lewis symbols or electron-dot symbols of the first twenty elements in the periodic table.
The electron dots in the Lewis structure are a convenient way to determine how many bonds an atom of an element can make.
For example, a hydrogen atom with one unaired dot can make one bond as in H-H. A bond is represented by a line between the
bonded atoms. A bond is formed by sharing unpaired valence electrons. It is called a covalent bond. Carbon, nitrogen, oxygen, and
fluorine with 4, 3, 2, and 1 unpaired dot can make 4, 3, 2, and 1 bond, e.g., in the following molecules: , , , and .
2.6.1 https://chem.libretexts.org/@go/page/372868
Atomic size
Electrons exist around the nucleus in a cloud-like appearance with no clearly defined boundaries. Therefore, the atomic size
generally refers to the covalent radius of an atom that is one-half of the distance between the nuclei covalently bonded in a
homonuclear molecule, like Cl2, I2, H2, as illustrated in Fig. 2.6.2.
Figure 2.6.2 : Diagram showing the atomic radius of H2. Source: modified from CK-12 Foundation / CC BY-SA
(https://creativecommons.org/licenses/by-sa/3.0)
Figure 2.6.3 : Illustration of the periodic trend of atomic sizes, Calculated Atomic Radii (in Picometers). Source:
https://chem.libretexts.org/@api/dek...jpg?revision=1
Ionization energy
The positively charged nucleus attracts the negatively charged electrons. Therefore removal of an electron from the atom requires
energy. The ionization produces a cation with fewer electrons than the parent neutral atom, i.e., cations.
Ionization energy
Ionization energy is the energy needed to remove an electron from a neutral atom, as in the following reaction.
1+ −
Na + ionization energy → Na +e
2.6.2 https://chem.libretexts.org/@go/page/372868
2. The ionization energy generally increases from left to right in a row because the valence electrons are in the same shell
while more protons add to the nucleus, which increases the pull on the valence electrons.
Figure 2.6.4 : The 1st ionization energy in kJ/mol. Source: Download for free at https://openstax.org/details/books/chemistry
Electronegativity
Definition: Electronegativity
Electronegativity is the ability of an atom in a compound to attract the bonding electron pair to itself. Electropositivity is the
opposite of electronegativity. Electronegativity is a property of an atom in a compound, i.e., a bonded atom, not a feature of an
individual atom.
There are several electronegativity scales. The most commonly used is the Pauling electronegativity scale. Fig. 2.6.5 shows the
electronegativities values on Pauling’s electronegativity scale.
2.6.3 https://chem.libretexts.org/@go/page/372868
Metallic character
The metallic character relates to the ease of losing an electron in a chemical reaction. The metallic character trend is opposite to the
trend of ionization energy.
Figure 2.6.6 : Summary of the periodic trend in the properties of elements. Source: https://www.hiclipart.com/free-
trans...ntzga/download
2.6: The periodic trends in properties of the elements is shared under a Public Domain license and was authored, remixed, and/or curated by
Muhammad Arif Malik.
2.6.4 https://chem.libretexts.org/@go/page/372868
CHAPTER OVERVIEW
3: Compounds
3.1: Bonding in compounds
3.2: Naming binary ionic compounds
3.3: Polyatomic ions and their compounds
3.4: Naming acids
3.5: Naming binary covalent compounds
3.6: Lewis structures of molecules
3.7: Molecular shapes –Valence shell electron pair repulsion (VSEPR) theory
3.8: Polarity of molecules
3.9: Intramolecular forces and intermolecular forces
3: Compounds is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
1
3.1: Bonding in compounds
Chemical bonds
What is a compound?
Compounds are a pure form of matter formed by atoms of more than one element combined in a constant whole number ratio.
The bonds connect the atoms in the compounds. Sharing or transferring some valance electrons from one atom to the other makes
the bonds. Noble gases have a full valence shell of eight valence electrons, except helium which has a full valence shell of two
valence electrons. The noble gases are the least reactive, i.e., the most inert group of elements.
Octet rule
The octet rule states that atoms of all elements other than noble gases tend to share, lose, or gain valence electrons to acquire
the electron configuration of the nearest noble gas having eight valence electrons.
Covalent bonds
A bond formed by sharing electrons is a covalent bond.
When a nonmetal atom combines with another nonmetal atom, they usually make a covalent bond. A covalent bond is a pair of
shared electrons, called a bonding pair of electrons, where each bonded atom contributes one electron. For example, chlorine has
seven valence electrons and needs one more to complete its octet. Hydrogen has one valence electron and requires one more to
acquire the electron configuration of helium, i.e., duet instead of the octet. Hydrogen and chlorine combine by sharing one electron
to make the compound HCl. Similarly, oxygen, nitrogen, and carbon make 2, 3, and 4 covalent bonds with hydrogen to complete
their octet and make compounds H2O, NH3, and CH4, respectively, as illustrated in Fig. 3.1.1.
Figure 3.1.1 : Examples of covalent bonds where Cl, O, N, and C acquire octet by sharing 1, 2, 3, and 4 electrons, respectively.
Hydrogen shares one electron and acquires a duet, i.e., the electron configuration of helium. 2nd row is the space-filling model of
the molecule where the white sphere is hydrogen, and the colored is the other atom. Source: "File:Kovalentne veze.png" by Drago
Karlo CC BY-SA 4.0
Two atoms can share one, two, or three electrons to make a single, a double, or a triple covalent bond. For example, H2 has a single
bond (H-H), O2 has a double bond (O=O), and N2 has a triple bond (N≡N), where each line between the atoms represent one
covalent bond. Fig. 3.1.2 illustrates the formation of three covalent bonds in N2. A valence electron pair that is not involved in
bonding is called a nonbonding pair. One nonbonding pair and three bonding pairs complete the octet of each nitrogen atom in the
N2 molecule, as shown in Fig. 3.1.2.
Note
The bonding pair of electrons counts towards the total valence electrons of each bonded atom., i.e., in H-H each hydrogen atom
has two valence electrons, and in :N≡N: each nitrogen has eight valence electrons; two in the nonbonding pair and six in three
bonding pairs.
3.1.1 https://chem.libretexts.org/@go/page/372888
Figure 3.1.2 : Illustration of three covalent bond formations between two nitrogen atoms. Source: File:Kovakentna veza
dušika.png" by Drago Karlo is licensed under CC BY-SA 4.0
Figure 3.1.3 : Quartz crystal composed of SiO2, and an illustration of covalent bonds in a giant molecule of SiO2. Source: Parent
Géry, Public domain, via Wikimedia Commons, and Roland Mattern / Public domain
Ionic bond
A bond formed by the transfer of electrons from one atom to the other atoms is an ionic bond.
A compound that has ionic bonds is an ionic compound. Usually, metal atoms lose electrons and become cations, and nonmetal
atoms gain electrons to become anions. The electrostatic attraction between the opposite charges holds the ions together in the ionic
compound. For example, sodium (Na) loses one electron, and fluorine (F) gains one electron to make a compound sodium fluoride
(NaF), as illustrated in Fig. 3.1.4.
Figure 3.1.4 : An example of ionic bond formation between Na and F. Source: modified from: Wdcf / CC BY-SA.
Ionic compounds
Table salt, i.e., NaCl is an example of an ionic compound. The Na completes its octet by losing one electron and becoming Na+
cation. Losing electrons reduces the electron-electron repulsion, but the electron-nucleus attraction remains the same.
Consequently, the electron cloud around the nucleus shrinks. Similarly, the chlorine atom has seven valence electrons. After
gaining one electron, it becomes Cl- anion with its octet complete. Gaining electrons increases the electron-electron repulsion, but
the electron-nucleus attraction remains the same. Consequently, the electron cloud around the nucleus expands. Fig. 3.1.5 illustrates
3.1.2 https://chem.libretexts.org/@go/page/372888
the formation of an ionic bond and the accompanying changes in the total number of electrons and sizes relative to the parent
neutral atoms in the case of NaCl formation.
Figure 3.1.5 : Illustration of an ionic compound formation by electron transfer from sodium (Na) to chlorine (Cl) making Na+
cation with one less electron and Cl- anion with one more electron than their parent neutral atoms. The Na+ cation is smaller than
Na, and the Cl- anion is large than Cl atom.
The ionic bond is not localized or unidirectional. The electrostatic force is all around the ions. Therefore, the cations surround the
anions, and the anion surrounds the cations in a regular array in a 3D crystal lattice. Fig 3.1.6 illustrates the structure of NaCl.
Note
1. The formula of the ionic compounds represents the simplest whole-number ratio of the atoms of the constituent elements.
2. A cation is always smaller in size than its parent neutral atom.
3. An anion is always larger than its parent neutral atom.
Figure 3.1.6 : Sodium Chloride from left: actual crystals in a water bubble within a 50 mm metal loop, model, and crystal lattice.
Na+ is purple or blue, and Cl- is green in the models. Source: Photograph by the NASA Expedition 6 crew, Public domain, via
Wikimedia Commons, Benjah-bmm27 (talk · contribs) / Public domain, and File: H Padleckas / Public domain
Usually, a metal and a nonmetal bond is ionic, and the bond between two nonmetals is covalent. Better criteria are based on the
difference in electronegativities of the bonded atoms. Electronegativity is the ability of an atom to attract a pair of bonded
electrons to itself. If the electronegativity difference is significant, the bonding electrons completely transfer to the more
electronegative atom, and the bond is ionic. There is no single value of electronegativity difference to separate ionic and
covalent bonds, but usually, the electronegativity difference of more than 1.8 results in an ionic bond. Otherwise, a covalent
bond, but the bonding electrons are more towards the more electronegative atom, making it a polar covalent bond. An
electronegativity difference less than 0.5 is considered a noncovalent bond, but a true noncovalent bond forms when the
bonded atoms are the same element.
Properties of compounds
The properties of the compounds are usually altogether different from the properties of their constituent elements. For example,
hydrogen (H2) is a gas that burns in oxygen, oxygen (O2) is a gas that assists combustion, but water (H2O) is a liquid that
extinguishes fire. Similarly, sodium (Na) is a soft metal that melts at 97.79 oC, chlorine (Cl2) is yellowish color gas, but sodium
chloride (NaCl) is a transparent crystal that melts at 801 oC.
3.1.3 https://chem.libretexts.org/@go/page/372888
The intermolecular interactions in covalent molecules are weak to moderate relative to the strength of covalent bonds or ionic
bonds. Therefore, the covalent molecules are usually gases like O2, NH3, CH4, liquids like H2O, or soft and low melting solids like
waxes, glucose (C6H12O6, mp 146oC).
Ions in the ionic compounds are held together by strong ionic bonds in a 3D array of crystal lattices. Therefore, ionic compounds
are usually hard solids with high melting points. For example, NaCl melts at 801 oC. Covalent compounds that exist as a 3D
network of covalent bonds, i.e., as giant molecules, are usually hard materials having a higher melting point than ionic compounds.
For example, SiO2 present in sand and quartz is a hard solid that melts at 1,710 oC. Diamond –the hardest known substance, is a
giant molecule of carbon atoms held together in a 3D network of covalent bonds that melts around 4,027 oC.
3.1: Bonding in compounds is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
3.1.4 https://chem.libretexts.org/@go/page/372888
3.2: Naming binary ionic compounds
Binary ionic compounds are compounds composed of monoatomic cations and monoatomic anions. For example, NaCl is a binary
ionic compound composed of monoatomic cations Na+ and monoatomic anions Cl-. Another example is CaCl2 composed of
monoatomic cations Ca2+ and monoatomic anions Cl-.
Example 3.2.1
Example 3.2.2
Calculate the charge on an iron ion in Fe2O3?
Solution
There are three oxygen anions, each with a -2 charge, making -6. So the total charge on two iron atoms should be +6, i.e., the
charge on iron atoms is +3. Answer: Fe3+.
Figure 3.2.1 : The charges on common ions. Download for free at https://openstax.org/details/books/chemistry.
3.2.1 https://chem.libretexts.org/@go/page/372890
Name of a monoatomic cation having a fixed charge
Alkali metals have +1, alkaline earth metals have +2, and aluminum has +3 charge. Their name is the name of the element ending
with ion. For example, Na+ is a sodium ion, Ca2+ is a calcium ion, and Al3+ is an aluminum ion.
The ions that are important in body fluids include sodium ion (Na+), potassium ion (K+), calcium ion (Ca2+), magnesium ion
(Mg2+), and chloride ion (Cl-), as shown in Fig. 3.2.2. Na+ is present in fluids inside the cells. It regulates and controls body
fluids. K+ is present in fluids outside the cells and regulates body fluids and cell functions. Ca2+ and Ma2+ are present in the
body fluids outside the cells, where Ca2+ is needed for muscle contraction, Mg2+ is needed for muscle contraction, nerve
control, and enzymes. Cl- is primarily present to balance the charge of the cations in the body fluids.
Figure 3.2.3 : Rules to write the formula of the binary ionic compounds.
3.2.2 https://chem.libretexts.org/@go/page/372890
Writing the names of an ionic compound from the formula
Writing the name of ionic compounds of cations with fixed charge
If the cation has a fixed charge in compounds, the name starts with the name of the element of the cation, followed by the name of
the anion without the world ion at the end. For example, KI is potassium iodide, and CaCl2 is calcium chloride.
Silver, zinc, and cadmium cations have fixed changes: Ag+, Zn2+, and Cd2+. Names of these cations are the names of the
element with or without charge shown in roman numerals, both ways it is correct.
Table 1: Examples of writing names from the formulae of binary ionic compounds
Example# Formula Name of the cation Name of the anion Name of the compound
3.2: Naming binary ionic compounds is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif
Malik.
3.2.3 https://chem.libretexts.org/@go/page/372890
3.3: Polyatomic ions and their compounds
Polyatomic ions
Polyatomic ions are molecular ions composed of two or more atoms bonded by covalent bonds and acting as a single unit, but
unlike molecules, they have a net charge on them.
The examples include cations like ammonium ion (NH ), and hydronium ion (H O ); and anions like hydroxide ion (OH ), and
+
4 3
+ −
cyanide ion (CN ). Usually, the name of polyatomic cations ends with –ium, and the name of polyatomic anions end with –ide,
−
except for oxyanions that have separate rules for their nomenclature.
Oxyanion
The oxyanions are oxides of nonmetals that are molecular ions. Examples include carbonate (CO ) , nitrate (NO ), phosphate ( 2−
3
2−
3
PO
3−
4
), and sulfate (SO ). The following guidelines will help remember the names and charges of oxyanions in most cases.
2−
4
valence elections in the central atom -2 x number of oxygen atoms = 5-(2x4) = -3; in (SO ), sulfur has 6 valence electrons
2−
and there are 4 oxygen atoms, so the charge = 6-(2x4) = -2; in (CO ), carbon has 4 valence electrons and there are 3
2−
3
oxygen atoms, so the charge = 4-(2x3) = -2; and in (NO ), nitrogen 6 valence electrons and there are 3 oxygen atoms, so
−
3
3
2−
4
2−
Oxyanions of halogens
Oxyanins of chlorine, bromine, and iodine are also common oxyanions with the following in common.
1. They have -1 charge,
2. a halogen with four oxygen is named by adding prefix "per-" to the name of the halogen with last syllable replaced with -
ate, for example; (ClO ) is perchlorate, (BrO ) is perbromate, and (IO ) is periodate.
−
4
−
4
−
4
3. a halogen with three oxygen is named as name of the halogen with last syllable replaced with -ate, for example; (ClO ) is −
3
4. a halogen with two oxygen is named as name of the halogen with last syllable replaced with -ite; for example, (ClO ) is −
2
5. a halogen with one oxygen is named by adding prefix "hypo-" to the name of the halogen with last syllable replaced with -
ite; for example, (ClO ) is hypochlorite, (BrO ) is hypobromite, and (IO ) is hypoiodaite.
− − −
3
Acids of oxyanions
Oxyanions are acids when their charge in neutralized with protons (H ). Names of the acids are the names of oxyanions with -
+
ate replaced with -ic acid and -ite replaced with -ous acid. For example: (HNO ) is nitric acid and (HNO ) is nitrous acid; (
3 2
H PO ) is phosphoric acid and (H PO ) is phosphorous acid; and (H SO ) sulfuric acid and (H SO ) is sulfurus acid.
3 4 3 3 2 4 2 3
The prefixes "per-" and "hypo-" in the cases of oxyanions of halogens remain in the acid name. For example: (HClO ) is 4
perchloric acid; (HClO ) is chloric acid; (HClO ) is chlorous acid; and (HClO) is hypochlorous acid.
3 2
Oxyanions with one proton attached but charge one them not fully neutralized, i.e., they are still polyatomic anion are named
beginning with hydrogen and ending with the name of the oxyanion. For example: (HSO ) is hydrogen sulfate; (HSO ) is −
4
−
3
3.3.1 https://chem.libretexts.org/@go/page/372891
hydrogen sulfite; and (HPO 2−
4
) is hydrogen phosphate.
Oxyanions with two protons attached but charge one them not fully neutralized, i.e., they are still polyatomic anion are named
beginning with dihydrogen and ending with the name of the oxyanion. For example, (H PO ) is dihydrogen phosphate
2
−
fTwo oxyanions containing a transition metal as the central atom in common use as reagents are in chemistry are chromate (
CrO 4
) and permanganate (MnO ). Table 3.3.1 lists the formulas and names of some of the common polyatomic ions.
2− −
4
(NH ) +
4
Ammonium (MnO ) −
4
Permanganate
(H3
O
+
) Hydronium (BrO ) −
4
Perbromate
(HO ) −
Hydorxide (IO )
−
4
Periodate
(CN ) −
Cynide (CrO 2−
4
) Chromate
(CO 2−
3
) Carbonate (CO 2−
2
) Carbonite
(NO ) −
3
Nitrate (NO ) −
2
Nitrite
(PO 3−
4
) Phosphate (PO 3−
3
) Phosphite
(SO 2−
4
) Sulfate (SO 2−
3
) Sulfite
(HCO ) −
3
Hydrogen carbonate (ClO ) −
4
Perchlorate
(HSO ) −
4
Hydrogne sulfate (ClO ) −
3
Chlorate
(HPO 2−
4
) Hydrogenphosphate (ClO ) −
2
Chlorite
(H2
PO
−
4
) Dihydrogenphosphate (ClO ) −
Hypochlorite
3.3: Polyatomic ions and their compounds is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad
Arif Malik.
3.3.2 https://chem.libretexts.org/@go/page/372891
3.4: Naming acids
Acids donate produce protons and anions when dissolved in water. The naming is different for acids that have oxyanions and the
others.
Naming as acids
Add prefix hydro- to the name of anion and replace the last syllable from –ide to –ic acid. For example, HCl is hydrochloric acid,
HCN is hydrocyanic acid, HI is hydroiodic acid. The acid name is preferred when the compound acts as an acid, particularly when
it is in solution form in water.
3 3
; (NO ) is nitrite and (HNO ) is nitrous acid; (SO ) is sulfate and {H2SO4}\)) sulfuric acid; (ClO ) is perchlorate and (
−
2 2
2−
4
−
4
HClO ) is perchloric acid; and (ClO ) is hypochlorite and (HClO) is hypochlorous acid.
−
One of the commonly encountered oxyacid is acetic acid. Acetic acid is an organic acid with the formula CH3COOH, where the
last hydrogen attached with oxygen is the acidic proton, and the other three oxygen attached with carbon are not acidic. The anion
from acetic acid is called acetate ion that has a formula CH3COO- which may also be written as C2H3O2-.
3.4: Naming acids is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
3.4.1 https://chem.libretexts.org/@go/page/372892
3.5: Naming binary covalent compounds
What is a binary covalent compound?
Binary covalent compounds are atoms of two different elements held together by covalent bonds. Usually, they are composed of
nonmetals elements, e.g., laughing gas NO, acid rain causing gas SO2, etc.
Mono- 1
di- 2
Tri- 3
tetra- 4
penta- 5
hexa- 6
hepta- 7
octa- 8
nona- 9
deca- 10
3.5: Naming binary covalent compounds is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif
Malik.
3.5.1 https://chem.libretexts.org/@go/page/372893
3.6: Lewis structures of molecules
Lewis symbols of elements
Lewis symbol of element shows the symbol of element with valence electrons shown as dots placed on top, bottom, left, and right
sides of the symbol. Valence electrons up to four are shown as a single dot on either side of the symbol. The 5th, 6th, 7th, and 8th
valence electron dots are paired with any of the first four dots. For example, represent
hydrogen, beryllium, boron, carbon, nitrogen, fluorine, and neon with 1, 2, 3, 4, 5, 6, 7, and 10 valence electrons, respectively.
Lewis symbols of first twenty elements are shown in section 2.6, Table 1.
The covalent bond is usually represented by a single line between the bonded atoms, e.g., the H2 molecule shown in the above
equation is generally shown as H-H. An example is a reaction between hydrogen having one valence electron and carbon having
four valence electrons react to form CH4 molecule.
Similarly, hydrogen reacts with nitrogen, oxygen, and fluorine to form the following molecules: , , and . Each line in
these molecules represents a bonding electron pair, and the pair of dots represent valence electrons that are not involved in
bonding, called lone pair of electrons. The lone pair is usually omitted from the Lewis structure unless it is needed to emphasize
their presence for some reason.
Example 3.6.1
Draw the Lewis structures of CH4, PCl3, CO2, and HCN
Solution
Step 1: Add the valence electrons of all the molecules' atoms:
CH4 has 4 valence electrons in C, and 1 in each of the four H: = 4 + 1x4 = 8 valence electrons
PCl3 has 5 valence electros in P and 7 in each of the three Cl: = 5 + 7x3 = 26 valence electrons
CO2 has 4 valence electrons in C and 6 in each of the two O: = 4 + 6x2 = 16 valence electrons
HCN has 1 valence electron in H, 4 in C, and 5 in N: = 1 + 4 + 5 = 10 valence electrons
Step 2: Place the element symbol with more valances, i.e., having more unpaired dots in its Lewis structure, in the center and
the rest of the atoms on four sides:
Step 3: Draw a line between the outer atom and the central atom to represent a single covalent bond:
3.6.1 https://chem.libretexts.org/@go/page/372894
Step 4: Every single bond consumes two valence electrons. Subtract the total number of valence electrons consumed in all the
bonds from the total valence electrons initially present in step 1:
CH4: 4 bond = 8 electrons consumed. 8 – 8 = 0 electrons left
PCl3: 3 bond = 6 electrons consumed. 26 – 6 = 20 electrons left
CO2: 2 bond = 4 electrons consumed. 16 – 4 = 12 electrons left
HCN: 2 bond = 4 electrons consumed. 10 – 4 = 6 electrons left
Step 5: Distribute the remaining electrons as lone pairs, first to the outer atoms to complete their octet (duet in the case of
hydrogen) and then to the central atom to complete its octet
Step 6: Check that the octet of each atom is complete (duet for hydrogen). If yes, the Lewis structure is complete, e.g., as in the
cases of CH4 and PCl3 in the present examples. If not, move one of the lone pair of electrons from a neighboring atom to make
a double bond, as shown by the red color arrows in the figure in the previous step. If the octet is still not complete, move one
more lone pair of electrons from a neighboring atom: from the same atom to make a triple bond, as in the case of HCN above,
or from another neighboring atom to make two double bonds, as in the case of CO2 above. The result is Lewis structures
shown below.
Note
To draw the Lewis structure of the most stable form, try to keep covalent bonds with an atom equal to the number of unpaired
dots in the Lewis symbol of the atoms. For example, have 1, 1, 2, 3, and 4 unpaired electrons. So,
hydrogen makes one, fluorine one, oxygen two, nitrogen 3, and carbon four covalent bonds in the stable molecules. When
pulling a lone pair from neighboring atoms is needed to result in a double or triple bond, it is preferable to keep the resulting
number of covalent bonds equal to the number of unpaired electrons in the Lewis symbol of the atom. For example, in the case
of CO2 molecule, if both lone pairs were pulled from the same oxygen in step 6 above, the resulting Lewis structure would
have been , which is technically correct Lewis structure but not the most stable form of this molecule. Note that the
latter structure has one covalent bond in one oxygen and three covalent bonds in the other instead of two covalent bonds
needed for oxygen.
Structure H:H
3.6.2 https://chem.libretexts.org/@go/page/372894
Phosphorous
Name Hydrogen Beryllium dichloride Boron trifluoride Sulfuric acid
pentafluoride
Valence electrons on
2 4 6 10 12
the central atom
3.6: Lewis structures of molecules is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
3.6.3 https://chem.libretexts.org/@go/page/372894
3.7: Molecular shapes –Valence shell electron pair repulsion (VSEPR) theory
The Lewis structure tells the connection between atoms and any lone pair present, but it does not tell the exact angles of bonds
around the central atom or the actual shape of the molecule. The conventional way of presenting a Lewis structure of a molecule
shows it as a planer, e.g., CH4 as which implies the molecule to be planar and H-C-H bond angles to be either 90o or 180o.
The actual CH4 molecule is nonplanar as with all H-C-H bond angles 109.5o. the following theory helps explain the actual
shapes of molecules.
Electron groups
A lone pair, a single bond, a double bond, and a triple bond, each of these is one electron group. This is because two elections of
single bond, four electrons of a double bond, and six electrons of a triple bond are located in the region along the axis of the bond,
i.e., they are grouped together. Similarly, a lone pair is located in a defined space around the atom. For example, carbon in methane
( ) has four electron groups that are the four single bonds (C-H bonds) around the carbon. Carbon in carbon dioxide (O=C=O)
has two electron groups that are the two double bonds around the carbon, and in H-C≡N has two electron groups that are a single
bond (C-H bond) and a triple bond (C≡N bond).
One electron group between two atoms is always a linear molecule. For example, H-H ,O=O, N≡N, and H-Cl , are linear
molecules, where hydrogen is white and chlorine is green in the H-Cl model.
Two-electron groups
Two-electron groups are farthest apart in a linear geometry with the central atom in the middle of the line and the bond angles of
180o around the central atom. The examples include CO2 and HCN , where the central carbon atom is gray,
hydrogen is white, nitrogen is blue, and oxygen is red.
Three-electron groups
Three-electron groups are farthest apart when they are at the corners of a triangle in a planar trigonal geometry with the central
atom in the middle of the triangle and the bond angles of 120o around the central atom. Examples include BF3 , and H2CO
, where boron is pink, F is green, carbon is gray, oxygen is red, and hydrogen is white.
If one of the electron domains is a lone pair, the electron domain geometry remains the same, but the geometry of the atoms in the
molecule, i.e., molecule geometry, is bent. For example, has three electron domains and trigonal planar electron domain
geometry, but there is one lone pair. So, the molecule geometry is bent as , where sulfur is yellow, and oxygen is red (lone
pair in not shown).
Four-electron groups
Four-electron groups are farthest apart when they are at the corners of a tetrahedron in a tetrahedral geometry with the central
atom at the center of the tetrahedron and the bond angles of 109.5o around the central atom as: . An example is methane CH4
3.7.1 https://chem.libretexts.org/@go/page/372895
, where carbon is gray, and hydrogens are white.
If one of the electron domains is a lone pair, the electron domain geometry is still tetrahedral, but the molecule geometry is
trigonal pyramidal as with three pereferal atoms at the corners of the triangel and the central atom raised to the top of the
pyramid. An example is ammonia (:NH3) , where nitrogen is blue, and hydrogens are white.
If two electron domains are lone pairs, the electron domains geometry is still tetrahedral, but the molecule geometry is bent. An
1 0 Linear Linear -
HCl
CO2
H2CO
CH4
:NH3
4 1 Tetrahedral Trigonal pyramidal 109.5o clipboard_e279a05d0785cb1c2a0
b2c4c0be750b03.png
3.7: Molecular shapes –Valence shell electron pair repulsion (VSEPR) theory is shared under a Public Domain license and was authored, remixed,
and/or curated by Muhammad Arif Malik.
3.7.2 https://chem.libretexts.org/@go/page/372895
3.8: Polarity of molecules
The polarity of an atom
The negative charge of electrons balances the positive charge of protons in an atom. The electrons symmetrically distributed around
the nucleus leave no negative or positive end. The atoms are nonpolar. Fig. 3.8.1 illustrates the polarity of a hydrogen atom with
color codes.
Figure 3.8.1 : Electrostatic potential map of hydrogen atom, shown nonpolar by green color code. Source: drawn using free
software https://chemagic.org/molecules/amini.html
Figure 3.8.2 : H-F molecule showing δ+ end in blue, δ- end in red and neutral part in green. Source: drawn using free software
http://molview.org/?cid=14917
Bond polarity is a vector that has a magnitude and direction and can be represented by an arrow, like other vectors, as shown in Fig.
3.8.3 for the case of a water molecule.
3.8.1 https://chem.libretexts.org/@go/page/372896
Figure 3.8.3 : The water molecule is made up of oxygen and hydrogen, with respective electronegativities of 3.44 and 2.20. The
electronegativity difference polarizes each H–O bond, shifting its electrons towards the oxygen (illustrated by red arrows). These
effects add as vectors to make the overall molecule polar. Source: Riccardo Rovinetti / CC BY-SA
(https://creativecommons.org/licenses/by-sa/3.0)
A bond is categorized as a nonpolar covalent, polar covalent, or an ionic bond based on the following convention: nonpolar
covalent if the electronegative difference of the bonded atom is less than 0.5, a polar covalent if the electronegativity difference is
between 0.5 to 1.9, and an ionic if the electronegativity difference is more than 1.9.
Figure 3.8.4 : Methane (CH4) with no polar bond is nonpolar. Source: drawn using free software
https://chemagic.org/molecules/amini.html
If there is only one polar bond in a molecule, then the molecule is polar, e.g., the H-F molecule shown in Fig. 3.8.2.
Figure 3.8.5 : Examples of symmetric molecules with all polar bonds but the molecule as a whole is nonpolar. The examples are
(from left to right) carbon dioxide CO2, boron trifluoride BF3, and carbon tetrafluoride CF4.
If there is more than one polar bond in a molecule, the molecule may be polar or may be nonpolar, depending on the symmetry of
the molecule: a) Polarity vector of individual bonds cancel out in symmetric molecules making the molecule nonpolar. For
example, symmetric molecules, like CO2, BF3, and CCl4, are nonpolar, although each bond in them is polar. Fig. 3.8.5 illustrates
the symmetric molecules that have polar bonds, but the polarity of bonds cancels each other, making the molecule nonpolar. b) If a
3.8.2 https://chem.libretexts.org/@go/page/372896
molecule has polar bonds and it is not symmetric, the polarity vectors do not cancel out, and the molecule is polar. Examples of
polar molecules include CHCl3, NH3, and H2O, as illustrated in Fig. 3.8.6.
Figure 3.8.6 : Non-symmetric molecules with polar bonds are polar molecules with a net δ+ end in red color and δ- ends in blue
color. The examples are (from left to right) chloroform (CHCl3), ammonia (NH3), and water (H2O).
3.8: Polarity of molecules is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
3.8.3 https://chem.libretexts.org/@go/page/372896
3.9: Intramolecular forces and intermolecular forces
There are electrostatic interaction between charges or partial charges, i.e., the same charges attract each other, and opposite charges
repel each other, as illustrated in Fig. 3.9.1.There are two types of electrostatic forces in compounds or molecules, intramolecular
forces that exist between the bonded atoms of a compound or a molecule, and intermolecular forces that exist between molecules as
described below.
Figure 3.9.1 : Like charges repel each other and opposite charges attract each other.
Intramolecular forces
Intramolecular forces are the chemical bonds holding the atoms together in the molecules. The three major types of chemical bonds
are the metallic bond, the ionic bond, and the covalent bond.
Metallic bond
Metals exist as a collection of many atoms as +ions arranged in a well-defined 3D arrangement called crystal lattice with some of
the outermost electrons roaming around in the whole piece of the metal, forming a sea of electrons around the metal atoms, as
illustrated in Fig. 3.9.2. The attraction between +ions and the sea of free moving electrons is the metallic bond that holds the atoms
together in a piece of metal. The metallic bond is usually the strongest type of chemical bond.
Ionic bond
When the electronegativity difference between bonded atoms is large, i.e., more than 1.9 in most cases, the bonding electrons
completely transfer from a more electropositive atom to a more electronegative atom creating a cation and an anion, respectively.
There is the electrostatic interaction between cation and anion, i.e., the same charges attract each other, and opposite charges repel
each other, as illustrated in Fig. 3.9.1. The cations and anions orient themselves in a 3D crystal lattice in such a way that attractive
interactions maximize and the repulsive interactions minimize, as illustrated in Fig. 3.9.3. Ionic bonds are usually weaker than
metallic bonds but stronger there the other types of bonds.
Figure 3.9.3 : Illustration of an ionic bond: the ions in ionic crystals orient in such a way to optimize the attractive forces between
opposite charges. Part of one layer of a crystal lattice is shown in this illustration, the 3D crystal lattice is illustrated in Fig. 3.1.6.
Source: Eyal Bairey / Public domain
3.9.1 https://chem.libretexts.org/@go/page/372897
Covalent bond
When the electronegativity difference between bonded atoms is moderate to zero, i.e., usually less than 1.9, the bonding electrons
are shared between the bonded atoms, as illustrated in Fig. 3.9.4. The attractive force between the bonding electrons and the nuclei
is the covalent bond that holds the atoms together in the molecules. The covalent bond is usually weaker than the metallic and the
ionic bonds but much stronger than the intermolecular forces.
Figure 3.9.4 : Illustration of a covalent bond in fluorine molecule: Jacek FH, CC BY-SA 3.0
<https://creativecommons.org/licenses/by-sa/3.0>, via Wikimedia Commons
Figure 3.9.5 : van Arkel-Ketelaar Triangle: plot of the difference in electronegativity ( χ) versus the average electronegativity in a
Δ
bond (χ average ). The top corner is mostly ionic, the lower left is metallic, and the lower right is covalent.
Intermolecular forces
Intermolecular forces are the electrostatic interactions between molecules. The intermolecular forces are usually much weaker than
the intramolecular forces, but still, they play important role in determining the properties of the compounds. The major
intermolecular forces include dipole-dipole interaction, hydrogen bonding, and London dispersion forces.
3.9.2 https://chem.libretexts.org/@go/page/372897
Dipole-dipole interactions
Polar molecules have permanent dipoles, one end of the molecule is partial positive (δ+) and the other is partial negative (δ-). The
polar molecules have electrostatic interactions with each other through their δ+ and δ- ends called dipole-dipole interactions,
though these interactions are weaker than ionic bonds. The polar molecules orient in a way to maximize the attractive forces
between the opposite charges and minimize the repulsive forces between the same charges, as illustrated in Fig. 3.9.6.
Hydrogen bonds
Hydrogen bonding is a dipole-dipole interaction when the dipole is a hydrogen bond to O, N, or F, e.g. in water molecules as
illustrated in Fig. 3.9.7. Although hydrogen bond is a dipole-dipole interaction, it is distinguished from the usual dipole-dipole
interactions because of the following special features.
1. The electronegativity difference between H and O, N, or F is usually more than other polar bonds.
2. The charge density on hydrogen is higher than the δ+ ends of the rest of the dipoles because of the smaller size of hydrogen.
3. The δ+ Hydrogen can penetrate in less accessible spaces to interact with the δ- O, N, or F of the other molecule because of its
small size.
Figure 3.9.7 : Illustration of hydrogen bonds by dotted lines in water molecules. Red balls are oxygen and white balls are hydrogen
atoms. Source: User Qwerter at Czech Wikipedia: Qwerter. Transferred from cs.Wikipedia to Commons by sevela.p. Translated to
english by by Michal Maňas (User:snek01). Vectorized by Magasjukur2 / Public domain
A hydrogen bond is usually stronger than the usual dipole-dipole interactions. Hydrogen bonding is the most common and essential
intermolecular interaction in biomolecules. For example, two strands of DNA molecules are held together through hydrogen
bonding, as illustrated in Fig. 3.9.8. Proteins also acquire structural features needed for their functions mainly through hydrogen
bonding.
3.9.3 https://chem.libretexts.org/@go/page/372897
Figure 3.9.8 : Chemical structure of DNA, with a colored label identifying the four bases as well as the phosphate and deoxyribose
components of the backbone. Dotted lines are the hydrogen bonds holding the two strands of DNA together. Source: Madprime
(talk · contribs) / CC0
Figure 3.9.9 : Illustration of transient dipole-induced dipole, i.e., London dispersion forces.
3.9: Intramolecular forces and intermolecular forces is shared under a Public Domain license and was authored, remixed, and/or curated by
Muhammad Arif Malik.
3.9.4 https://chem.libretexts.org/@go/page/372897
CHAPTER OVERVIEW
4: Stoichiometry –the quantification of chemical reactions
4.1: Stoichiometry
4.2: The mole
4.3: Chemical reaction
4.4: Patterns of chemical reactions
4.5: Oxidation-reduction reactions
4.6: Energetics of chemical reactions
4.7: Stoichiometric calculations
4: Stoichiometry –the quantification of chemical reactions is shared under a Public Domain license and was authored, remixed, and/or curated by
Muhammad Arif Malik.
1
4.1: Stoichiometry
What is Stoichiometry
Stoichiometry (stoi·chi·om·e·try /ˌstɔɪkiˈɒmɪtri/) is the study of the quantities of substances and energy consumed and
produced in chemical reactions.
The basis of the stoichiometric calculations is the law of conservation of mass which states that the mass is neither created nor
destroyed in a chemical reaction. Another form of the law states that atoms are neither created nor destroyed in a chemical reaction.
It is the basis of stoichiometric calculations that are described in this chapter.
It can be concluded from the law of mass action that atoms of each element and their masses are the same in reactants and products.
A balanced chemical equation shows atoms of each element and the total mass of reactants equal to that in the product, as
illustrated in Fig. 4.1.1. The number of atoms and molecules is related to their quantity in moles through Avogadro's number. The
mole, in turn, is related to the mass of the substance through molar mass in grams. These relationships are described in the next
sections.
Atomic mass
Atomic mass is the weighted average of the masses of the isotopes present in a natural sample of the element, as explained with an
example calculation in section 2.3. The mass of a single atom or molecule is expressed in the atomic mass unit (amu), which is
equal to th of the mass of C isotope of carbon that is unbound.
1
2
12
6
Atomic mass is listed in a periodic table as a number below the symbol and name of the element, as illustrated in Fig. 4.1.2. Atomic
mass is listed as a number without a unit because it is the mass of an atom in amu and it is also the mass of one mole (molar mass)
of the atom in grams. Molar mass is often used in stoichiometry calculation as explained in the next sections.
Figure 4.1.2 : Copy and Paste Caption here. (Copyright; author via source)Atomic number and atomic mass in a periodic table for
Helium
4.1: Stoichiometry is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
4.1.1 https://chem.libretexts.org/@go/page/372939
4.2: The mole
Avogadro’s number and mole
Just like the dozen is equal to 12, the Avogadro's number 6.02214076 x 1032 is exact by definition, but usually, 6.022 x 1023 is used
in calculations with 4 significant figures.
Mole (mol)
The mole is a SI unit of the amount of a substance that is equal to 6.02214076 x 1023 particles of the substance.
The particle of a substance are usually atoms, ions, or molecules. For example, 6.02214076 x 1023 atoms of C isotope is one 12
6
mole of C. The number 6.02214076 x 1032 is exact by definition, but usually, 6.022 x 1023 is used with 4 significant figures.
12
6
One mole of a substance is equal to one Avogadro’s number of atoms, molecules, or formula units of the substance. i.e.,
23
1 Avogadro's number of particles = 1 mol = 6.022 × 10 particles
, where the first factor is used to convert a number of moles to a number of particles and the second for a number of particles to
number moles conversions, as explained in the following examples.
Example 4.2.1
Step 2. Write the two conversion factors from equality between the given and the desired quantity.
23
6.022 × 10 particles 1 mol
and
23
1 mol 6.022 × 10 particles
Step 3. Multiply the given quantity with the conversion factor that cancels the given unit and leaves the desired unit in the
answer.
23
6.022 × 10 molecules C9 H8 O4
21
0.0139 mol C9 H8 O4 × = 8.37 × 10 molecules C9 H8 O4
1 mol C 9 H8 O4
Example 4.2.2
How many moles of aspirin ((C 9
H O
8 4
) are in 9.50 x 1025 molecules of aspirin?
Solution
Step 1. Write the given quantity and the desired quantity.
4.2.1 https://chem.libretexts.org/@go/page/372941
Given: 9.50 x 1025 molecules of aspirin, Desired: ? mol of aspirin
Step 2. Write the two conversion factors from equality between the given and the desired quantity.
23
6.022 × 10 particles 1 mol
and
23
1 mol 6.022 × 10 particles
Step 3. Multiply the given quantity with the conversion factor that cancels the given unit and leaves the desired unit in the
answer.
1 mol C 9 H8 O4
25
9.50 × 10 molecules C9 H8 O4 × = 158 mol C 9 H8 O4
23
6.022 × 10 molecules C9 H8 O4
these equalities between a mole of a substance and the moles of the element in it gives two conversion factors for the calculations.
Example 4.2.3
Step 3. Multiply the given quantity with the conversion factor that cancels the given unit and leaves the desired unit in the
answer.
12 mol H
3.0 mol C 6 H12 O6 × = 36 mol H
1 mol C 6 H12 O6
Example 4.2.4
Solution
Step 1. Write the given quantity and the desired quantity.
Given: 0.50 moles of Ca(NO 3
) , Desired: ? mol of O
2
Step 2. Write the two conversion factors from equality between the given and the desired quantity. Note that NO is −
polyatomic ion that has three oxygen atoms in it. There are two NO in the formula unit as shown by subscript 2 outside the
−
bracket enclosing the polyatomic anion (\ce{NO3^-}\). So, the equality is:
1 mole Ca(NO 3
)
2
= 6 mole O,
and the two conversion factors from the equality are:
1 mol Ca(NO ) 6 mol O
3 2
and
6 mol O 1 mol Ca(NO )
3 2
Step 3. Multiply the given quantity with the conversion factor that cancels the given unit and leaves the desired unit in the
answer.
4.2.2 https://chem.libretexts.org/@go/page/372941
6 mol O
0.50 mol Ca(NO ) × = 3.0 mol O
3 2
1 mol Ca(NO )
3 2
Molar mass
g
The mass of one mole of a substance, in mol
, is called the molar mass.
Recall that one mole = 1 Avogadro's number, i.e., 6.022×1023 atoms, molecules, or formula units of a substance. Fig. 4.2.1 helps in
visualizing the molar masses of aluminum, copper, and carbon. The molar mass of an element or a compound is a reasonable
quantity to be measured on an analytical balance commonly available in laboratories, while the mass of an individual atom or
molecule is too small to be easily measured. That is why the mole is commonly used in stoichiometric calculations.
Figure 4.2.1 : A mole of aluminum, copper, and carbon visualized. Source: photo: R. Press/NIST; graphic design: N.
Hanacek/NIST
compound, take the atomic masses of the constituent elements from a periodic table, multiply them with the number of atoms of the
element in the formula of the compound, and then add these numbers, as explained in the following examples. Note that the unit
can also be written as: g.mol-1.
g
mol
4.2.3 https://chem.libretexts.org/@go/page/372941
Example 4.2.5
Calculate the molar mass of NaOH?
Solution
Step 1. Find the atomic masses of the constituent elements from the periodic table.
Na = 22.990 g.mol-1, O = 15.999 g.mol-1, H = 1.008 g.mol-1.
Step 2. Multiply the atomic masses with the number of atoms in the fourmula.
1x22.990 g.mol-1 Na, 1x15.999 g.mol-1O, 1x1.008 g.mol-1 H.
Step 3. Add all the numbers from step 2.
1x22.990 g.mol-1 Na + 1x15.999 g.mol-1O + 1x1.008 g.mol-1 H = 39.997 g.mol-1 NaOH
Example 4.2.6
g
is a conversion factor converting the mass of a substance in grams to the amount of the
substance in moles. The conversions are explained in the following examples.
Example 4.2.7
18.02 g H2 O 1 mol H2 O
and
1 mol H2 O 18.02 g H2 O
Step 4. Multiply the given quantity with the conversion factor that cancels the given unit
and leaves the desired unit in the answer.
4.2.4 https://chem.libretexts.org/@go/page/372941
18.02 g H2 O
2.50 mol H2 O × = 45.1 g H2 O
1 mol H2 O
Example 4.2.8
How many moles are present in 2.50 g of aspirin? (Aspirin C 9
H O
8 4
, molar mass 180.2 g.mol-1 )
Solution
Step 1. Write the given quantity and the desired quantity.
Given: 2.50 g C 9
H O
8 4
, Molar mass of C
9
H O
8 4
= 180 g.mol-1, Desired: ? g H2O
Step 2. Calculate the molar mass of the substance.: given 180.2 g.mol-1
Step 3. Write the molar mass of the substance, and it's reciprocal as the two conversion factors.
180 g C H O 1 mol C H O
9 8 4 9 8 4
and
1 mol C H O 180 g C H O
9 8 4 9 8 4
Step 4. Multiply the given quantity with the conversion factor that cancels the given unit
and leaves the desired unit in the answer.
1 mol C H O
9 8 4
2.5 g C H O × = 0.0193 mol C H O
9 8 4 9 8 4
180 g C H O
9 8 4
Example 4.2.9
Step 4. Multiply the given quantity with the conversion factor that cancels the given unit
and leaves the desired unit in the answer.
1 mol N aOH
10.0 g N aOH × = 0.250 mol NaOH
40.00 g N aOH
4.2: The mole is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
4.2.5 https://chem.libretexts.org/@go/page/372941
4.3: Chemical reaction
What is a chemical reaction?
A chemical reaction is a combination, separation, or rearrangement of atoms in a substance.
Atoms are healed together in a substance by chemical bonds. The combination makes bonds, separation breaks bonds, and
rearrangement breaks some of the old bonds and makes some new bonds in the substances during a chemical reaction. It results in
new substances with a different composition of elements than the starting substances. The starting substances are called reactants
and the new substances formed are called products.
When a chemical bond is broken energy is required and when a chemical bond is formed energy is released. The amount of energy
depended on the chemical bond, but for the same bond, the energy that needs to break is the same energy released to make the
bond. Therefore, energy is released in some chemical reactions and absorbed in others.
Examples of chemical reactions are photosynthesis which converts carbon dioxide and water into glucose in the green leaves of
plants using energy from sunlight. Digestion of food is a chemical reaction that releases the energy needed for the functioning of
living things. The burning of a candle is a chemical reaction that converts the organic compounds in the fule to carbon dioxide,
water, and heat energy. Rusting iron is another chemical reaction that converts the element iron to compound iron oxide.
Figure 4.3.1 : Examples of indications of chemical reactions: from left a) butane burning with flame and light, b) Silver chloride
formation with precipitation, c) copper ammonia complex formation with a color change to greenish to dark blue, and d) Reaction
of limestone with HCl with carbon dioxide gas formation. Source: a) Renate90 / CC BY-SA
(https://creativecommons.org/licenses/by-sa/3.0) , b) Luisbrudna / CC BY-SA (https://creativecommons.org/licenses/by-sa/4.0), c)
https://www.youtube.com/watch?v=7TIIOj-SHkw&t=102s, d) Alessandro e Damiano / CC BY
(https://creativecommons.org/licenses/by/4.0)
Chemical equaitron
1. The formulas of the reactants are written on the left side separated by a + sign.
2. Formulas of the products are written on the right-side separated by a + sign.
3. An arrow pointing in the direction of products separates the reactants from the products.
4.3.1 https://chem.libretexts.org/@go/page/372942
Fig. 4.3.2 illustrates the photosynthesis reaction between carbon dioxide (CO and water (H 2 2
O producing glucose (C 6
H
12
O
6
and
oxygen (O :2
sunlight
6 CO +6 H O−−−−−→C H O +6 O
2 2 6 12 6 2
For example, Fig. 4.3.2 shows coefficients 6 for carbon dioxide, water, and oxygen, but no coefficient for glucose means the
coefficient is actually one for glucose in a balanced chemical equation for photosynthesis reaction.
Figure 4.3.2 : Model and components of a balanced chemical equation with the example of photosynthesis reaction.
For example, combustion of butane (ceC 4H 10 in lighter shown in Fig. 4.3.1. can be represented as:
, where all the reactants and products are in the gas phase. Reaction of NaCl and AgNO3 in the water that results in a precipitate, as
shown in Fig. 4.3.1 is:
, where NaCl, AgNO3, and NaNO3 are in the water, and AgCl precipitates out as white solid. The reaction of limestone with HCl,
shown in Fig. 4.3.1 is:
4.3.2 https://chem.libretexts.org/@go/page/372942
Reaction conditions, catalysts, or heat can be written above or below the arrow. Heat can also be represented by ∆ symbol, e.g.,
Δ
2 NO (g) −
↽⇀
− N O (g)
2 2 4
H +O ⟶ H O
2 2 2
Note that hydrogen is written as H2 (not H) and oxygen as O2 (not O) because these reactants usually exist as molecules, not as
atoms.
The next step is to add coefficients to balance atoms of each element on the two sides of the equation. For example, in the above
equation hydrogen is balanced but oxygen is not. Balance oxygen by changing the coefficient of water from 1 to 2:
H +O ⟶ 2H O
2 2 2
Caution
Subscripts in the formulae can not be changed, as they are constant. For example, if O2 is changed to O in the above equation
to balance the oxygen, it is incorrect as O2 is molecular oxygen which is a different chemical than atomic oxygen O.
The coefficient is a multiplier of each subscript in the formula, i.e., in 2H2O there are 2x2 = 4 hydrogen and 2x1 =2 oxygen. Now
look for the other elements again: note that hydrogen atoms have changed to 4 on the right side. To balance hydrogen, change the
coefficient of H2 from 1 to 2:
2H +O ⟶ 2H O
2 2 2
Check again: Now, the atoms of each element are the same on both sides, i.e., the equation is balanced, as illustrated in Fig. 4.3.3.
Figure 4.3.3 : Illustration of a balanced chemical equation of water formation from hydrogen and oxygen. Atoms of each type are
balanced. Source: Kvr.lohith / CC BY-SA (https://creativecommons.org/licenses/by-sa/4.0)
1. if an element occurs only in one reactant and one product, balance it first,
2. if there is a free element in the reactants or products, balanced it last,
3. if there is a polyatomic ion in reactants and products, balance the number of polyatomic ions as a unit, i.e., add coefficient
to but do not change the subscripts in the polyatomic ion or of any formula.
4. in some cases, a fractional coefficient is needed to balance the equation, in that case first use the fractional coefficient and
when the equation is balanced, remove the fraction by multiplying all the coefficients with a common multiplier,
5. if the set of coefficients in the balanced equation is not in the simplest whole-number ratio, the equation is still balanced,
but it is recommended to convert the set of coefficients to the simplest whole-number ratio.
4.3.3 https://chem.libretexts.org/@go/page/372942
The following examples explain the guidelines.
Figure 4.3.4 : Propane torch being used to solder copper pipes for residential water mains. Source: Kevin L Neff / CC BY
(https://creativecommons.org/licenses/by/2.0)
Example 4.3.1
Combustion of propane (C3H8) is used in gas welding, as shown in Fig. 4.3.4. The propane reacts with oxygen and produces
carbon dioxide and water. Write and balance the chemical equation of the reaction?
Solution
Start with writing the correct formulae of reactants and products in a chemical equation form:
C H +O ⟶ CO +H O
3 8 2 2 2
According to rule 1, look at carbon or hydrogen first. Both are not balanced. Balance the carbon by changing the coefficient of
CO2 from 1 to 3:
C H +O ⟶ 3 CO +H O
3 8 2 2 2
C H +O ⟶ 3 CO +4 H O
3 8 2 2 2
Now, look at oxygen (rule 2) and balance it by changing the coefficient of O2 from 1 to 5:
C H +5 O ⟶ 3 CO +4 H O
3 8 2 2 2
Figure 4.3.5 : Illustration of a balanced chemical equation of propane combustion reaction. Source: Kvr.lohith / CC BY-SA
(https://creativecommons.org/licenses/by-sa/4.0)
Example 4.3.2
The reaction between Aluminium metal powder and iron oxide, called thermite reaction, is highly exothermic that melts the
ion. The thermite reaction is used for railroad welding as shown in Fig. 4.3.6. Wright and balance the reaction equation.
Solution
Start with writing the correct formulae of reactants and products in a chemical equation form:
Al + Fe O ⟶ Fe + Al O
2 3 2 3
4.3.4 https://chem.libretexts.org/@go/page/372942
Rule 1 does not apply as all elements are in one reactant and one product. According to rule 2, look at either one of the
compounds. Balance aluminum by changing the coefficient in the eactant from 1 to 2:
2 Al + Fe O ⟶ Fe + Al O
2 3 2 3
2 Al + Fe O ⟶ 2 Fe + Al O
2 3 2 3
All atoms are balanced at this stage and coefficients are in the simplest whole-number ratio. So the equation is balanced.
Figure 4.3.1 : Thermite reaction proceeding. Shortly after this, the liquid iron flows into the mold around the rail gap. Source:
http://en.Wikipedia.org/wiki/Image:V...ewelding-1.jpg, Public domaine.
Example 4.3.3
Methanol (CH4O) reacts with oxygen and produces carbon dioxide and water. Write and balance the chemical reaction
equation?
Solution
The starting equation with correct formulas is:
CH O + O ⟶ CO +H O
4 2 2 2
According to rule 1, look at carbon and hydrogen first. Carbon is already balanced. There are 4 hydrogens on the left but only
two on the right side, so balance them by changing the coefficient of water from 1 to 2:
CH O + O ⟶ CO +2 H O
4 2 2 2
Now, look at oxygen (rule 2). There are three oxygen atoms on the right (one in methanol and two in free element oxygen) but
four oxygen on the right, it is not balanced. Changing the coefficient of CH4O from 1 to 2 balances the oxygen, but carbon and
hydrogen go out of balance. Changing the coefficient of O2 from 1 to 2 does not balance the oxygen. A fractional coefficient
3/2 for O2 works at this stage (rule 4):
3
CH O + O ⟶ CO +2 H O
4 2 2 2 2
All elements are balanced, i.e., the equation is balanced, but, according to rule 5, it is better to remove the fraction by
multiplying each coefficient in the equation with 2:
2 CH O + 3 O ⟶ 2 CO +4 H O
4 2 2 2
The set of coefficients in the balanced equation is already in the simplest whole-number ratio, so no further step is needed in
this case.
4.3.5 https://chem.libretexts.org/@go/page/372942
Example 4.3.4
Aluminum reacts sulfuric acid to produce aluminum sulfate and hydrogen gas. Wright a balanced chemical equation for the
reaction.
Solution
The starting equation with correct formulas of reactants and products is:
Al + H SO ⟶ Al (SO ) +H
2 4 2 4 3 2
All elements occur in one reactant and one product, so rule 1 do not apply. According to rule 2, leave aluminum till the end
and, according to rule 3, balance the polyatomic ion (SO42-) as a unit by changing the coefficient of sulfuric acid from 1 to 3:
Al + 3 H SO ⟶ Al (SO ) +H
2 4 2 4 3 2
Al + 3 H SO ⟶ Al (SO ) +3 H
2 4 2 4 3 2
Double-check at the end –All elements are balanced, and all the coefficients are in the simplest whole-number ratio. No further
step is needed.
4.3: Chemical reaction is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
4.3.6 https://chem.libretexts.org/@go/page/372942
4.4: Patterns of chemical reactions
General types of chemical reactions
There are several ways to classify chemical reactions. The general types of chemical reactions fall in the categories of combination,
decomposition, replacement, and combustion reactions, as illustrated in Fig. 4.4.1.
Figure 4.4.1 : Illustration of four basic patterns of chemical reactions: combination, decomposition, single & double replacement,
and combustion reactions. Source: https://www.hiclipart.com/free-trans...dunqu/download
Combination reactions
A compound is synthesized or formed from two or more substances, e.g.:
C +O ⟶ CO
2 2
2H +O ⟶ 2H O
2 2 2
2 Mg + O ⟶ 2 MgO
2
2 Na + Cl ⟶ 2 NaCl
2
CaO + CO ⟶ CaCO
2 3
Fig. 4.4.2 shows an example of a hydrogen with oxygen combination reaction that is being developed for use as a fuel in the future.
Figure 4.4.2 : Berlin public bus using hydrogen as a fuel in an internal combustion engine (ICE). Hydrogen is stored in ten pressure
cylinders of 50 kg of hydrogen at 350 bar. Reaction: \{\ce{2H2 + O2 -> 2H2O}\) + heat. Source: StralsundByzantion / CC BY-SA
(https://creativecommons.org/licenses/by-sa/3.0)
Decomposition reactions
The decomposition reactions are the reverse of the combination reaction, i.e., one compound splits apart into two or more
substances, usually by heating, e.g.:
H CO ⟶ H O + CO
2 3 2 2
4.4.1 https://chem.libretexts.org/@go/page/372943
Δ
CaCO −
→ CaO + CO
3 2
2 KClO −
→ 2 KCl + 3 O
3 2
Electrolyisis
Fig. 4.4.3 illustrates the last reaction, i.e., decomposition of water through electrolysis.
El ectrol yisis
Fig. 4.4.4 shows an example of a single replacement reaction of magnesium resulting in hydrogen gas formation.
4.4.2 https://chem.libretexts.org/@go/page/372943
Figure 4.4.4 : Single replacement
reaction of magnesium with hydrochloric acid producing hydrogen gas:
Mg(s) + 2 HCl(aq) ⟶ MgCl(aq) + H (g)↑ . Source: taken form
2
NCSSM online youtube video, 05/18/20,
https://www.youtube.com/watch?v=OBdgeJFzSec
Double replacement reactions or metathesis involve the mutual exchange of partners between two substances, e.g. the following
precipitation reactions:
Combustion reactions
Combustion is a reaction of a substance with oxygen, often with the formation of flame and release of much heat, e.g.:
C H + 12 O ⟶ 8 CO + 8 H O + Heat
8 16 2 2 2
C +O ⟶ CO + Heat
2 2
2H +O ⟶ 2 H O + Heat
2 2 2
2 Mg + O ⟶ 2 MgO + Heat
2
Figure 4.4.5 : An image of Ignited Magnesium burning in the air in normal conditions. Yannickcosta1, CC BY SA3,
https://commons.wikimedia.org/wiki/File:Magburn1.jpg
Usually, combustion is considered as the reaction of a substance containing carbon and hydrogen with oxygen resulting in carbon
dioxide, water, flame, and heat, e.g., burning methane on a kitchen stove:
CH +3 O ⟶ CO + 2 H O + Heat
4 2 2 2
4.4.3 https://chem.libretexts.org/@go/page/372943
Oxidation-reduction reactions
The oxidation-reduction or redox reaction involves the exchange of electrons. For example, reactions between a metal and
nonmetal involve the transfer of electrons from the metal to the nonmetal forming an ionic bond, as shown in Fig. 4.4.6.
Acid-base reactions
The acid-base reactions involve the transfer of protons from an acid to a base, as shown in Fig. 4.4.7.
Figure 4.4.7 : Proton transfer from HCl to NH3 is an example of an acid-base reaction: HCl + NH 3
⟶ NH Cl
4
. Hydrochloric
acid (in beaker) reacts with ammonia fumes to produce ammonium chloride (white smoke). Source:
Walkerma, Public domain
Precipitation reactions
These are double displacement reactions in water that results in the precipitation of one of the products, as shown in Fig. 4.4.8.
Figure 4.4.8 : Lead (II) iodide precipitates when potassium iodide is mixed with lead (II) nitrate:
Pb ( NO ) (aq) + 2 NaI(aq) ⟶ 2 NaNO (aq) + PbI (s)↓
3 2 3 2
. Source: PRHaney / CC BY-SA
(https://creativecommons.org/licenses/by-sa/3.0)
The precipitation reactions and the acid-base reactions are described in the later chapters. The oxidation-reduction reactions are
discussed in the following section.
4.4: Patterns of chemical reactions is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
4.4.4 https://chem.libretexts.org/@go/page/372943
4.5: Oxidation-reduction reactions
What are Oxidation and Reduction?
The oxidation-reduction is defined in three ways:
1. Oxidation is the loss of electrons and reduction is the gain of electrons. The word OIL RIG helps in remembering this
definition.
2. The addition of oxygen is oxidation, and the removal of oxygen is a reduction.
3. The removal of hydrogen is oxidation, and the addition of hydrogen is a reduction.
The oxidation-reduction or in short redox reaction is one of the most common types of chemical reactions happening in and around
us. For example, rusting of metals, photosynthesis, digestion of food, and combustion of fuels are redox reactions.
Figure 4.5.1 : Green patina on the statue of liberty is a result of the oxidation of copper. Source: https://www.hiclipart.com/free-
trans...hybox/download
The transfer of electrons becomes apparent when the reaction is split into the oxidation-half and reduction-half. The oxidation-half
is:
4.5.1 https://chem.libretexts.org/@go/page/372944
2+ −
Cu ⟶ Cu +2 e
The electrons lost in an oxidation-half must be equal to the electrons gained by the accompanying reduction-half. Multiplying the
oxidation-half with 2 makes the electrons lost equal to the electrons gained in the reduction half in the above reaction. Then adding
the oxidation and the reduction half gives the overall reaction:
2+
2 Cu ⟶ 2 Cu + 4 e−
2−
O + 4 e− ⟶ 2 O
2
The chemical equations can be manipulated like algebraic equations, i.e., they can be multiplied or divided by a constant,
added, and subtracted, as demonstrated in the example of the copper redox-half reactions above. Note that electrons on one
side of the equation canceled the electrons on the other side of the equation during the addition operation.
A silver color zinc strip dipped in copper nitrate solution becomes coated with a layer of reddish color copper, as shown in Fig.
4.5.2. The molecular equation of the reaction, that shows formula units of compounds in the reactants and products, is:
The complete ionic equation for the reaction is obtained by showing the dissolved ionic compounds as ions, e.g.:
2+ − 2+ −
Zn(s) + Cu (aq) + 2 NO3 (aq) ⟶ Cu(s) + Zn (aq) + 2 NO3 (aq)
The net ionic equation that is the addition of the oxidation-half and the reduction-half is:
2+ 2+
Zn(s) + Cu (aq) ⟶ Cu(s) + Zn (aq)
Note that NO3- was on both sides of the complete ionic equation and has been canceled out in the net ionic equation:
2+ − 2+ −
Zn(s) + Cu (aq) + 2 NO (aq) ⟶ Cu(s) + Zn (aq) + 2 NO (aq)
3 3
Ions that do not participate in the chemical reaction are called spectator ions, and they appear on both sides of the molecular
equation, like NO3- in this case.
4.5.2 https://chem.libretexts.org/@go/page/372944
Biological oxidation and reduction reactions
In the redox reactions involving metal species, the transfer of electrons is usually evident through the oxidation-half and the
reduction-half reactions, as in the above examples. In the organic and biochemical redox reactions, the transfer of electrons is
usually not so obvious, but the transfer of hydrogen or oxygen is usually apparent. For example, the metabolism of methanol
H C−OH starts with oxidation through the loss of hydrogen:
3
H C−OH ⟶ H C=O + 2 H
3 2
H C=O ⟶ HCOOH
2
Finally, the formic acid (HCOOH) is oxidized by gaining oxygen and forming carbon dioxide and water:
2 HCOOH ⟶ 2 O=C=O + 2 H O
2
The increase in the C-O bond from a single bond in methanol to two double bonds (four C-O bonds) in carbon dioxide is a clear
indication of the oxidation of the carbon.
Removal of hydrogen is also oxidation, e.g., hydroxymalonate is oxidized to oxomalonate by an enzyme hydroxymalonate
dehydrogenase:
The reverse of these, i.e., removal of oxygen and addition of hydrogen is reduction.
4.5: Oxidation-reduction reactions is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
4.5.3 https://chem.libretexts.org/@go/page/372944
4.6: Energetics of chemical reactions
Energetics deals with the energies involved in chemical reactions. There are two types of energies: the first is related to
thermodynamics which is the energy released or absorbed when the reactants convert to the products, and the second is related to
the kinetics of the reaction, i.e., the energy that the reacting molecules must possess to surpass the energy barrier for conversion to
the products.
Figure 4.6.1 : A reaction coordinate diagram for the reaction: F + H C−I ⟶ F−CH + I . The green sphere is F-, purple is
−
3 3
−
I-, gray is C, and white is H. ∆G‡ is the activation energy, and ∆Go is the overall energy released or absorbed in a chemical reaction.
The starting point shows the energy level of reactants. Remember that every substance has internal energy which is manifested as
temperature. The molecules are moving around due to the internal energy and collied with other molecules. The collision between
the reactants is usually the first requirement for the reaction to proceed.
The F- ion in this example collides with the target H3C-I from the side opposite to the C-I bond. The collision at an angle other than
180o to the C-I bond is less effective or ineffective in initiating the C-I bond breaking. So, the proper orientation of the reacting
molecules at the time of the collision is the second requirement.
After a properly oriented collision, the C-I bond starts breaking, the F-C bond starts making, and the energy of the system rises to a
maximum level where the bond breaking and making are about halfway. The specie at the maxima of the energy curve is called
activated complex or a transition state. The difference in energy between the reactants and the transition state is called activation
energy (∆G ‡ ). The reactants must have enough kinetic energy to surpass the activation energy barrier, which is the third
requirement for the reaction to happen. A summary of the basic requirements of a chemical reaction to happen is the following.
The activated complex forms if the basic requirements are fulfilled. The activated complex rolls down the hill on the energy scale
and settles at the energy level of the products where the old bonds are entirely broken, and the new bonds are fully formed. The
graph of the reaction progress versus the energy, as in Fig. 4.7.1 is called the reaction coordinate diagram.
4.6.1 https://chem.libretexts.org/@go/page/372946
Exothermic and endothermic reactions
The energy absorbed or released in the form of heat at constant pressure conditions is called enthalpy change (∆H), which is almost
the same as the internal energy absorbed or released (∆G) in most of the reactions.
A chemical reaction that releases heat is exothermic. A chemical reaction that absorbs heat is an endothermic reaction.
Bond-forming is always exothermic, and bond-breaking is the opposite, i.e., endothermic, as illustrated in Fig. 4.7.2.
Figure 4.6.2 : Bond forming releases heat, i.e., exothermic, and breaking the same bond absorbed the same energy, i.e.,
endothermic. Source: https://www.hiclipart.com/free-trans...alsen/download
In a chemical reaction, some bonds break, and some bonds form. If the bond-breaking absorbs less heat than the heat released in
the new bond making, the reaction is exothermic. The products are lower in energy than the reactants, and the ∆H is negative for
exothermic reactions. The opposite is true for an endothermic reaction, i.e., the bond-breaking absorbs more heat than the bond-
forming; the products are higher in energy than the reactants, and ∆H is positive, as illustrated in Fig. 4.8.3.
Figure 4.6.3 : Reaction Coordinate Diagrams showing exothermic with products lower in energy than the reactants (left and middle)
and endothermic with products higher in energy than the reactant (right) reaction. It also shows that the reaction on the left is the
slowest due to the highest activation energy (∆G‡), and the reaction in the middle is the fastest due to the lowest activation energy
(∆G‡). Source: AimNature / CC BY-SA (https://creativecommons.org/licenses/by-sa/3.0)
An example of an exothermic reaction is the combustion of methane that releases heat used for cooking.
o
CH +2 O ⟶ CO +2 H O ΔH =−891 kJ
4 2 2 2
Note that the sign of ∆H is –ve for exothermic and +ve for an endothermic reaction. Sometimes, the energy is shown as a product
for exothermic reaction, and as reactant for edothermic reaction. For example, the above two creations may be written as:
CH +2 O ⟶ CO + 2 H O + 891 kJ
4 2 2 2
6 CO + 6 H O + 2800 kJ ⟶ C H O +6 O
2 2 6 12 6 2
4.6.2 https://chem.libretexts.org/@go/page/372946
Figure 4.6.4 : Cold Hot Pack used for medication. Source: Mamun2a, CC BY-SA 4.0 <https://creativecommons.org/licenses/by-
sa/4.0>, via Wikimedia Commons
The instant hot or cold packs used in hospitals are based on the heat of dissolution of salts in water, as shown in Fig. 4.7.4. The salt
and the water are in separate pockets in the pack. When the seal is ruptured, the salt dissolves in water and releases or absorbs heat.
The dissolution of CaCl2 salt in water is an exothermic process that is the source of heat in a hot pack:
o
CaCl (s) ⟶ CaCl (aq) ΔH =−82 kJ
2 2
The dissolution of NH4NO3 salt in water is an endothermic process that is utilized to absorb heat in a cold-pack.
o
NH NO (s) ⟶ NH NO (aq) ΔH =−26 kJ
4 3 4 3
Any factor that increases the rate of collisions, enhances the proper orientation or increases the kinetic energy of the molecules
causes an increase in the rate of reaction. The factors include concentration, temperature, and catalysts.
In the breathing process, oxygen (O2) binds with hemoglobin (Hb) in the lungs.
Hb(aq) + O2(g) → HbO2(aq)
Patients having breathing problems are given breathing masks with a higher concentration of oxygen than in the atmosphere to
increase the rate of oxygen binding with the hemoglobin.
Figure 4.6.5 : Illustration of higher the concentration, the more frequent the collisions leading to faster reaction. Source: Sadi
Carnot / Public domain
Effect of temperature
The kinetic energy of molecules at a given temperature follows Boltzman distribution, as illustrated in Fig. 4.8.6. Increased
temperature increases the average kinetic energy of the molecules which increases the fraction of molecules with more than the
activation energy.
4.6.3 https://chem.libretexts.org/@go/page/372946
An increase In temperature increases the rate of chemical reactions—generally, a 10 oC increase in temperature doubles the
rate of a chemical reaction.
Figure 4.6.6 : Illustrating increased temperature increases the average kinetic energy of the molecules that increases the fraction of
molecules with more than the activation energy barrier (∆G‡). Source: modified from “TDF / CC0”
Some practical examples of the use of this principle are the following. Storing foods in refrigerators at lower temperatures
decreases the rate of reactions resulting in longer life of the foods. Cooking foods in a pressure cooker increases the temperature
resulting in faster cooking than in an open pan. In some cardiac surgeries, the body temperature is lowered to 28 oC to decrease the
rate of metabolism which decreases the oxygen demand so that the heart may be stopped temporarily for the surgery.
The catalysts and enzymes increase the rate of reaction by decreasing the energy of activation of the reaction through an
alternate route, as illustrated in Fig. 4.7.7.
The catalysts and the enzymes do not consume in the reaction -they regenerate and repeat the action. Enzymes are the catalysts in
biochemical reactions. Enzymes also increase the rate of reaction by binding with the reactants and properly orienting them for the
reaction.
Figure 4.6.1 : Effect of catalysts –When the catalyst is added, the activation energy ∆G ‡ decreases. However, the energy absorbed
or released, i.e., ∆Go remains constant. Source: Bkell, Public domain.
4.6: Energetics of chemical reactions is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif
Malik.
4.6.4 https://chem.libretexts.org/@go/page/372946
4.7: Stoichiometric calculations
Conversion factors from a chemical equation
Stoichiometry pronounced as “stɔɪkiˈɒmɪtri” is the calculation of the amount of reactants and products in a chemical reaction. It is
based on the fact that a balanced chemical equation is also a set of mole-to-mole equalities between the reactants and the products.
Each equality gives two conversion factors that allow calculating the mole of one substance from the given mole of any other
substance in the equation.
Fig. 4.6.1 lists the chemical equation for photosynthesis reaction, the mole-to-mole equalities from the equation, and the two
conversion factors from each of the equality, as an example. The conversion factors are used to calculate the unknown quantity in
the mole from the known quantity in the mole of any other reactant or product in the same chemical equation, as explained in the
following examples.
Figure 4.7.1 : Balanced chemical equation of the photosynthesis reaction, mol-to-mol equalities, and conversion factors derived
from the equation.
Example 4.7.1
Calculate the moles of glucose produced from 3.0 moles of carbon dioxide in the photosynthesis reaction?
Solution
i. Given: 3.0 mole CO2, Desired: ? moles of C6H12O6
ii. Chemical equation: 6 CO 2
+6 H O ⟶ C H
2 6 12
O
6
+6 O
2
v. Calculations:
4.7.1 https://chem.libretexts.org/@go/page/372945
1 mol C 6 H12 O6
3.0 mol CO2 × = 0.50 mol C 6 H12 O6
6 mol CO2
Example 4.7.2
Magnesium reacts with HCl by this reaction: Mg(s) + 2 HCl(aq) ⟶ MgCl (aq) + H (g)↑
2 2
. Calculate the moles of Mg
needed to produce 3 moles of H2?
Solution
i. Given: 3 mol of H2. Desired: ? Moles of Mg.
ii. The chemical equation is given.
iii. The desired equality: 1 mol Mg = 1 mol H2.
1 mol Mg
iv. The desired conversion factor: 1 mol H 2
v. Calculation:
1 mol Mg
3 mol H 2 × = 3 mol M g
1 mol H 2
Example 4.7.3
Calculate grams of AgCl precipitate formed from 2.0 moles of CaCl2 consumed in the following reaction:
Solution
i. Given: 2.0 moles of CaCl2. Desired: ? g AgCl
ii. Molar mass of AgCl = 1x107.87 g Ag.mol-1 + 1x35.45 g C.mol-1 = 143.3 g AgCl.mol-1
iii. The chemical equation is given.
iv. The desired equality: 1 mol CaCl2 = 2 mol AgCl, and 1 mol AgCl =143.3 g AgCl.
v. The desired conversion factors:
2 mol AgCl 143.3 g AgCl
and
1 mol CaCl2 1 mol AgCl
vi. Calculation:
2 mol AgCl
143.3 g AgCl
2.0 mol CaCl2 × × = 573.3 g AgCl
1 mol CaCl2 1 mol AgCl
Example 4.7.4
How many grams of carbon dioxide are needed to react with 2 moles of water in the photosynthesis reaction?
Solution
i. Given: 2.0 moles of H2O Desired: ? g CO2.
4.7.2 https://chem.libretexts.org/@go/page/372945
ii. Molar mass of CO2 = 1x12.011 g C.mol-1 + 2x15.999 g O.mol-1 = 44.009 g CO2.mol-1
iii. The chemical equation: : 6CO2 + 6 H2O → C6H12O6 + 6O2.
iv. The desired equality: 6 mol CO2 = 6 mol H2O, and 1 mol CO2 = = 44.009 g CO2.
v. The desired conversion factors:
6 mol CO2 44.009 g CO2
O and
6 mol H 2 1 mol CO2
vi. Calculation:
6 mol CO2 44.009 g CO2
2.0 mol H2 O × × = 88 g CO2
6 mol H 2 O 1 mol CO2
Mass of given substance⇒mole of the given substance⇒mole of the desired substance⇒grams of the desired substance.
The reciprocal molar mass of the given substance is the first conversion factor, the mole to mole conversion factor from the
chemical equation is the second conversion factor, and the molar mass of the desired substance is the third conversion factor
needed. Make sure that each conversion factor cancels the denominator unit of its multiplier to the right, and the desired unit is left
in the answer. The following examples explain these calculations.
Example 4.7.5
How many grams of Mg are needed to produce 1.01 g of H2 gas in this reaction:
Solution
i. Given: 1.01 g H2 Desired: ? g Mg.
ii. Molar mass of H2 = 2 x1.008 g H.mol-1 = 2.016 g H2.mol-1, and molar mass = 24.305 g Mg.mol-1.
iii. The chemical equation is given in the problem.
iv. The desired equalities : 1 mol Mg = 1 mol H2, 1 mol H2 = 2.016g H2, 1 mol Mg = 24.305 g Mg
v. Calculate by multiplying the given quantity consecutively with the three desired conversion factors from the equalities:
1 mol H 2 1 mol Mg
24.305 g Mg
1.01 g H2 × × × = 11.9 g Mg
2.016 g H 2 1 mol H 2 1 mol Mg
Example 4.7.6
How many grams of glucose are produced if 22.0 g of carbon dioxide is consumed in the photosynthesis reaction?
Solution
i. Given: 22.0 g CO2 Desired: ? g C6H12O6.
ii. Molar masses: of CO2 = 1x12.011 g C.mol-1 + 2x15.999 g O.mol-1 = 44.009 g CO2.mol-1, and molar mass of C6H12O6 =
6x12.011 g C.mol-1 + 12x1.008 g H.mol-1 + 6x15.999 g O.mol-1 = 180.156 g C6H12O6.mol-1 .
iii. The chemical equation: 6 CO 2
+6 H O ⟶ C H
2 6 12
O
6
+6 O
2
iv. The desired equalities: 6 mol CO2 = 1 mol C6H12O6, 1 mol CO2 = 44.009 g CO2, 1 mol C6H12O6 = 180.156 g C6H12O6.
4.7.3 https://chem.libretexts.org/@go/page/372945
v. Calculate by multiplying the given quantity consecutively with the three desired conversion factors from the equalities:
Fig. 4.6.2 illustrates what each conversion factor does in the case of the above example number 4.6.5
Figure 4.7.2 : Illustration of the operation of conversion factors in a mass-to-mass calculation based on a balanced chemical
equation of example 4..6.6 for question: "calculate grams of C6H12O6 produced from 22.0 g CO2 in photosynthesis reaction
6 CO + 6 H O ⟶ C H O + 6 O ”.
2 2 6 12 6 2
4.7: Stoichiometric calculations is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
4.7.4 https://chem.libretexts.org/@go/page/372945
CHAPTER OVERVIEW
5: Solutions
5.1: Introduction to solution
5.2: Solubility
5.3: Electrolytes
5.4: Concentration of solutions
5.5: Osmosis
5: Solutions is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
1
5.1: Introduction to solution
Solutions are all around us, e.g., air, seawater, body fluids, metal alloys are solutions. Fig. 5.1.1 illustrates that air is a mixture of
nitrogen, oxygen, carbon dioxide, and some other gases; Fig. 5.1.2 illustrates that seawater is a mixture of water, chloride, sodium,
sulfate, magnesium, and some other ions, and Fig. 5.1.3. illustrates that about 60% of the human body is composed of solutions
called body fluids.
Figure 5.1.1 : Air is a solution of nitrogen (N2), oxygen (O2), carbon dioxide (CO2), and some other gases . Source:
RoRo / CC0
Figure 5.1.2 : Seawater is a solution of water and several ions, expressed in wt/wt%. Source: derivative work: Tcncv (talk)Sea_salt-
e_hg.svg: Hannes Grobe, Alfred Wegener Institute for Polar and Marine Research, Bremerhaven, Germany; SVG version by Stefan
Majewsky / CC BY-SA (https://creativecommons.org/licenses/by-sa/2.5)
Figure 5.1.3 : This schematic shows the relative volumes of the different fluid compartments in an adult male. Source: Alan Sved
and David Walsh / CC BY-SA (https://creativecommons.org/licenses/by-sa/4.0)
What is a solution?
A solution is a homogeneous mixture of two or more pure substances.
The substance that is in a large amount in the solution is called the solvent.
The substance that is in smaller amounts in a solution is called the solute.
5.1.1 https://chem.libretexts.org/@go/page/372207
For example, the air is a solution in which nitrogen is the solvent, and water is the solvent in seawater and body fluids. Oxygen,
carbon dioxide, and water vapors are solutes in the air; and sodium, chloride, sulfate, magnesium, and some other ions are solutes
in seawater.
Types of solution
The solutions are generally classified in two ways: i) based on the physical state of the solution and the solute, and ii) based on the
particle size of the solute.
Types of solution based on the physical state of the solution and the solute
The solutions can be classified based on the physical state of the solution, solvent, and solute. For example, the air is gas in a gas
solution; carbonated water is a gas in a liquid solution; vinegar is a liquid in a liquid solution; metal alloys are solid in solid
solutions. Table 5.1.1 lists the major types of solutions, solvents, and primary solutes in them.
Table 1: Examples of main types of solutions and solvent and major solute in them.
Type Example solvent Primary solute
Solution
A solution is a homogeneous mixture comprising smaller component/s called solute/s of small molecules or ions comparable in
size to the molecules of a larger component called the solvent.
For example, NaCl dissolved in water is a solution. The solute is almost uniformly distributed in the solvent, making a
homogeneous mixture. The solute does not separate by filtration or by a semipermeable membrane but can be separated by some
other physical process. For example, the distillation process separates a solid in a liquid or a liquid into a liquid solution. The
solution is transparent, though it may be colored. A light passing through a solution is not visible, as shown in Fig. 5.1.4.
Figure 5.1.4 : Demonstration of Tyndall effect: a red-laser bean passing through a solution is not visible (left), but visible when
passing through a colloid (right). Source: https://youtu.be/8Xcpq6e8pBY, Creative Commons Attribution license (reuse allowed)
Suspension
A suspension is a heterogeneous mixture of solvent and solute particles of larger than 10,000 Å.
For example, muddy water is a suspension. If the suspension is allowed to stand, the suspended particles settle down and separate.
The suspended particles can be filtered out. Some medicines, e.g., milk of magnesia, are suspensions. It is instructed to shake just
before administering medicine to re-suspend the settled suspension.
5.1.2 https://chem.libretexts.org/@go/page/372207
Colloid
A colloid falls between a solution and a suspension. The colloidal particles are larger molecules like proteins or groups of
molecules or ions.
Unlike a suspension, the colloids usually do not settle if allowed to stand. The colloidal particles can not be filtered but can be
separated by a semipermeable membrane. When a light beam passes through a colloid, it scatters by the colloid particles, called the
Tyndall effect, and becomes visible, as shown in Fig. 5.1.4.
Examples of colloids include:
1. fog and clouds that are liquid water droplets dispersed in air;
2. smoke that is solid carbon particles dispersed in air;
3. whipped cream that is air dispersed in a liquid;
4. styrofoam is a gas dispersed in a solid; and
5. ager medium that is liquid dispersed in a solid medium.
6.
Figure 5.1.5 : the image of the earth showing saltwater in the sea, fresh water on the Antarctic ice sheet, and clouds. Source: by
NASA mission AS17. Public domain.
5.1.3 https://chem.libretexts.org/@go/page/372207
Figure 5.1.6 : A model of water molecule showing oxygen in red and hydrogen in white color balls overplayed with an electrostatic
potential map showing the partial positive (+d) region in red and the partial negative (-d) region in blue, drawn using a free web
app https://molview.org
When water mixes with other polar substances, like ethanol, some of the hydrogen bonding between water molecules replace
with similar hydrogen bonding with ethanol molecules. Since the electrostatic potential energy is similar, the natural tendency
to go towards more dispersion drives the dispersion of ethanol molecules uniformly in water resulting in the solution.
Figure 5.1.7 : Models showing hydrogen bonding in ethanol (left), water (middle), and ethanol water solution (right). Source:
modified from Benjah-bmm27, and snek01. Public domain
Ionic compounds are held together by electrostatic forces between opposite ions, i.e., ionic bonds. When an ionic compound is
added to water, water molecules surround the cation and establish ion-dipole interaction by orienting their partial -ve end to the
cation. Similarly, water molecules establish ion-dipole interaction with anions by orienting their partial +ev end towards the anion,
as illustrated in Fig. 5.1.8
The ion-dipole interactions, along with nature’s tendency to disperse the particles, are usually strong enough to overcome the
ionic bonds, dissociate the compounds into ions, and disperse them almost uniformly in the water.
5.1.4 https://chem.libretexts.org/@go/page/372207
Figure 5.1.8 : Dissolution of sodium chloride in water through dissociation into ions and dispersion in water driven by ion-dipole
interactions and nature’s tendency to disperse particles. Source: Andy Schmitz / CC BY
(https://creativecommons.org/licenses/by/3.0)
The separation of the cations from the anions of the ionic compound is called dissociation.
The formation of a layer of water molecules around ions, driven by ion-dipole interactions, is called hydration.
Non-polar substances, like vegetable oil or gasoline, do not dissolve in water. The molecules in non-polar substances have only
London dispersion forces. They easily dissolve in non-polar solvents like hexane or carbon tetrachloride that have similar London
dispersion forces among their molecules.
The fact that ionic and polar substances dissolve in polar solvents and non-polar substances dissolve in non-polar solvents of
similar intermolecular interactions is called “like dissolves like.”
5.1: Introduction to solution is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
5.1.5 https://chem.libretexts.org/@go/page/372207
5.2: Solubility
Solubility and its related terminologies
The majority of solutes do not dissolve in water or other solvents in all proportions.
What is solubility
The maximum proportion of the solute that can dissolve in a given amount of the solvent, usually expressed in grams of solute
in 100 grams of solvent, is called the solubility of the solute in the solvent.
1. Substances that make a solution when mixed in any proportion are called miscible. For example, ethanol and water are
miscible.
2. Some substances make a solution when mixed in some proportion but not in all proportions; these are called partially
miscible. For example, n-butanol can mix in water up to 7.3 g n-butanol/100 ml water.
3. Substances that almost do not dissolve in each other are called immiscible. For example, n-Hexane is immiscible in water.
4. The solution that has not yet reached its solubility limits and can dissolve more solute added is called an unsaturated
solution.
5. The solution that has reached its solubility limits and can not dissolve if more solute is added to it is called a saturated
solution.
When a solute is added to a saturated solution, it does dissolve, but, at the same time, the dissociated components recombine to
form the crystals of the solute, i.e., recrystallize, at the same rate, so that there is no net dissolution, as illustrated in Fig. 5.2.1.
Figure 5.2.1 : Illustration of unsaturated and saturated solution and the dynamic equilibrium between dissolution and
recrystallization in the saturated solution. Source: No machine-readable author provided. Romary assumed (based on copyright
claims). / CC BY-SA (http://creativecommons.org/licenses/by-sa/3.0/)
For example, when a saturated solution of sugar in water is heated, it can dissolve more sugar. Fig 5.2.2 show the solubility vs
temperature curves for some compounds.
5.2.1 https://chem.libretexts.org/@go/page/372208
Figure 5.2.2 : Solubility of some ionic and polar molecular compounds in water as a function of temperature. Download for free at
https://openstax.org/details/books/chemistry.
When a hot saturated solution is cooled, the concentration of solute in the solution becomes above the solubility limits, making
a supersaturated solution.
The supersaturated solution is unstable and ultimately crystallizes out the excess solute leaving any impurities in the solution, as
illustrated in Fig. 5.2.3. This process is called re-crystallization, which is used to purify the solutes.
The crystallization of excess solute from a supersaturated solution is responsible for some medical problems like gout and
kidney stone. Gout is the crystallization of uric acid in the cartilage, tendons, and soft tissues when the concentration of the uric
acid in blood plasms exceeds its solubility limit of ~7 mg/100 mL at 37 oC. It causes redness, swelling, and pain in the affected
area, as illustrated in Fig. 5.2.4. Kidney stones are solid materials formed in the urinary tubes, as illustrated in Fig. 5.2.5.
Kidney stones are the result of the crystallization of excess calcium phosphate, calcium oxalate, or uric acid in the urine.
5.2.2 https://chem.libretexts.org/@go/page/372208
Figure 5.2.4 : Illustration of gout as a result of the recrystallization of uric acid in the cartilage, tendons, and soft, causes swelling,
joint stiffness, and aching around the joint in the foot. Source: www.scientificanimations.com/ / CC BY-SA
(https://creativecommons.org/licenses/by-sa/4.0)
Figure 5.2.5 : Kidney Stones illustrated. Source: Blausen.com staff (2014). "Medical gallery of Blausen Medical 2014".
WikiJournal of Medicine 1 (2). DOI:10.15347/wjm/2014.010. ISSN 2002-4436/ CC BY
(https://creativecommons.org/licenses/by/3.0)
Opposite to the solids and liquids, the solubility of gases generally decreases with an increase in temperature.
That is why the carbonated water releases dissolved gas when heated, causing pressure increase, which, in turn, causes the bursting
of the soda can
Henry's law
The solubility of gases in liquids is directly proportional to the pressure of the gas above the liquid.
An increase in pressure causes a decrease in the gas volume that increases the gas concentration. More frequent collision of the gas
molecules with the gas-liquid boundary in a concentrated solution causes an increase in the rate of dissolution of the gas in the
liquid, as illustrated in Fig 5.2.6. The opposite happens when the gas pressure decreases. For example, carbon dioxide starts
bubbling out when a soda can is open because the gas escapes resulting in a decrease in the gas pressure above the liquid and a
decrease in the solubility of the gas in water.
Figure 5.2.6 : The Solubility of a gas increases as the partial pressure increases at a constant temperature. Source: Abozenadah, H.,
Bishop, A., Bittner, S. and Flatt, P.M. (2017) Preparatory Chemistry. CC BY-NC-SA. Available at:
https://wou.edu/chemistry/courses/on...ory-chemistry/
5.2.3 https://chem.libretexts.org/@go/page/372208
Solubility guidelines for dissolution of ionic compounds in water
If the solubility of a compound is less than 0.01 mol/L, it is considered insoluble.
The solubility of ionic compounds in water depends on the nature of the compound. For example, lead(II)iodide (PbI2) and silver
chloride (AgCl) are insoluble in water because the solubility of PbI2 is 0.0016 mol/L of the solution and the solubility of AgCl is
about 1.3 x 10-5 mol/L of solution. Potassium iodide (KI) and Pb(NO3)2 are soluble in water. When aqueous solutions of KI and
Pb(NO3)2 are mixed, the concentration of PbI2 in the mixture goes above its solubility limits, and it precipitates out, as illustrated in
Fig. 5.2.7.
(https://creativecommons.org/licenses/by-sa/3.0)]
There are no fail-proof guidelines for predicting the solubility of ionic compounds in water. However, the following guideline can
predict the solubility of most ionic compounds.
Soluble ions
1. Salts of alkali metals (Li1+, Na1+, K1+, Rb1+, Cs1+ ) and ammonia (NH41+) are soluble. No exceptions.
2. Salts of nitrate (NO31-), acetate (CH3COO1-), and perchlorate (ClO41-) are soluble. No exceptions.
3. Salts of chloride (Cl1-), bromide (Br1-), and Iodide (l1-) are soluble, except when the cation is Pb2+, Hg22-, or Ag1+.
(Remember the acronym “LMS” based on the first letter of the element name, or phrase ‘Let Me See” to recall Lead,
Mercury, and Silver.)
4. Sulfates (SO42-) are soluble except when the cation Pb2+, Hg22-, Ag1+, or a heavy alkaline earth metal ions: calcium (Ca2+),
barium (Ba2+), or strontium (Sr2+). ((recall “Let Me See” for Lead, Mercury, and Silver. Remember the acronym “CBS”
based on the first letter of the element name, or the phrase “Come By Soon” to recall calcium, barium, and strontium.)
Insoluble ions
1. Hydroxide (OH1-) and sulfides (S2-) are insoluble except when the cation is an alkali metal, ammonia, or a heavy alkaline
earth metal ions: Ca2+, Ba2+, and Sr2+. (Recall the phrase “Come By Soon” to recall calcium, barium, and strontium.)
2. Carbonates (CO32-), phosphates (PO43-), and oxide (O2-) are insoluble except when the cation is an alkali metal, ammonia
3. If there is a conflict between the two guidelines, then the guideline listed first has priority. For example, the salts of
insoluble ions become soluble when the cation is an alkali metal, ammonia (rule#1).
5.2.4 https://chem.libretexts.org/@go/page/372208
Figure 5.2.8 : The precipitates of some insoluble ionic compounds formed by mixing the aqueous solution of appropriate soluble
ionic compounds. The precipitates are from the left: white Calcium sulfate (CaSO4), black Iron(II) hydroxide (Fe(OH)2, brown
Iron(III) hydroxide (Fe(OH)3), and blue Copper(II) hydroxide (Cu(OH)2). Source: https://youtu.be/jltLlzZ6FqU
Fig. 5.2.8 shows precipitates of some insoluble ionic compounds formed by mixing aqueous solutions of appropriate soluble ionic
compounds. The precipitation can be predicted, as illustrated in Fig. 5.2.9. List the ions of the soluble ionic compounds and then
cross-combine the cations of one with the anion of the other to make the potential products. If any of the potential products is an
insoluble ionic compound, it precipitates out.
Figure 5.2.9 : Cross-combine the cation-anion in the reactants. If any of the cross-combination is an insoluble salt, it will precipitate
out, e.g: NaOH(aq) + MgCl (aq) → Mg(OH) ( s) ↓ +NaCl(aq).
2 2
5.2: Solubility is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
5.2.5 https://chem.libretexts.org/@go/page/372208
5.3: Electrolytes
What is an electrolyte?
Electric current is defined as the movement of electric charges. The substances through which an electric current can flow are
called electrical conductors, and the others are electrical nonconductors. Metals are electrical conductors because valence electrons
of metal atoms can move around in a piece of metal. Ionic compounds are composed of cations and anions, but the ions in a solid
can not move around. Therefore, solid ionic compounds are electrical nonconductors. Pure water does not have a sufficient
concentration of ions in it and is an electrical nonconductor. The ionic compounds dissociate into ions when dissolved in water. The
solution of ionic compounds in water is an electrical conductor because the ions can move around in the solution, as illustrated in
Fig. 5.3.1.
Figure 5.3.1 : Pure water (A) and solid ionic compounds (B) are electrical noncondutors, but the solution of ionic compounds in
water is an electrical conductor. Source: https://www.hiclipart.com/free-trans...cmeih/download
Substances that produce electrically conducting solution when dissolved in water (or in another polar solvent) are called
electrolytes.
All ionic compounds, acids, and bases produce ions in water and are classified are electrolytes.
Substances that produce an electrically nonconducting solution when dissolved in water are called nonelectrolytes.
Molecular compounds other than acids and bases, such as methanol, acetone, sugar, and glucose, remain neutral molecules when
dissolved in water. The molecular solutes, other than acids and bases, are nonelectrolytes.
5.3.1 https://chem.libretexts.org/@go/page/372210
A strong electrolyte does not mean that it is necessarily highly soluble in water. It means that the portion of the solute that
dissolves, it also dissociates 100% into ions in water, e.g., all ionic compounds. The solubility Ca(OH)2 is only 0.16 g Ca(OH)2/100
g water at 20 oC, but all the dissolved Ca(OH)2 dissociates into Ca2+ and OH- ions.
Strong bases are hydroxides of alkali metals, i.e., LiOH, NaOH, KOH, RbOH, and CsOH, and hydroxides of heavy alkaline earth
metals, i.e., Ca(OH)2, Sr(OH)2 and Ba(OH)2, which are strong electrolytes. Strong acids, i.e., HCl, HBr, HI, HClO4, HNO3, and
H2SO4, are molecular compounds but are strong electrolytes because they dissociate almost 100% into ions when dissolved in
water. For example, HCl almost wholly dissociates into ions when dissolved in water.
1+ 1−
HCl(g) + H2 O(l) → H3 O (aq) + Cl (aq)
Substances that partially dissociate into ions when dissolved in water are weak electrolytes. Weak acids and weak bases are
weak electrolytes.
Week acids and week bases, like acetic acid (CH3COOH) and ammonia (NH3), are soluble in water, but partially dissociate into
ions. For example, if 1-mole acetic acid or 1-mole of ammonia is dissolved in 1 liter of water at room temperature, they establish
the following equilibrium between dissolved molecules and dissociated ions:
Water
1− 1+
CH3 COOH(l) ⟶ CH3 COOH(aq) + H2 O(l) ⟵⃗ CH COO (aq) + H3 O (aq),
3
Water
1−
NH3 ( g) ⟶ NH3 (aq) + H2 O(l) ⟵⃗ NH 1+ (aq) + OH (aq),
4
where only about 0.4% of the dissolved molecules dissociate into ions, the remaining about 99.6% of molecules remain neutral.
Weak acids and weak bases are weak electrolytes.
Fig. 5.3.2 illustrates the difference between nonelectrolytes, strong electrolytes, and weak electrolytes.
Figure 5.3.2: Illustration of a nonelectrolyte that does not conduct electricity, a strong electrolyte that has high electrical
conductivity, and a weak electrolyte that has low electrical conductivity. Source: Karishma50/
(https;//creativecommons.org/licences/by-sa/4.0)
Equivalent
The amount of molecules and atoms is usually measured in moles. Ionic compounds are composed of ions but are overall neutral
because the +ve charge is balanced by the –ve charge. Therefore, the mole ratio of cations to anions is not always one to one. For
example, NaCl has a one-to-one mole ratio of Na+ and Cl-, but CaCl2 has a one-to-two ratio of Ca2+ and Cl-. A new unit, called
equivalent (officially abbreviated as Equiv but commonly abbreviated as Eq), is introduced, to differentiate between a mole of ion
and a mole of charge on the ion.
5.3.2 https://chem.libretexts.org/@go/page/372210
The equivalent is the amount of a substance needed to react with or supply one mole of hydrogen ions (H+) in an acid-base
reaction, or react with or supply one mole of electrons in a redox reaction. In other words, the number of equivalents of a given
ion in a solution is equal to the number of moles of that ion multiplied by its valence.
+ + 2+ 2+ 3+ 3+
1 mol Na = 1 Eq Na , but 1 mol Ca = 2 Eq Ca , and 1 mol Al = 3 Ea Al
Similarly,
− − 2− 2− 3− 3−
1 mol Cl = 1 Eq Cl , but 1 mol CO3 = 2 Eq CO3 and 1 mol PO 4 = 3 Ea PO 4
The solutions of electrolytes are overall electrically neutral, i.e., the number of equivalents of cations is equal to the number of
equivalents of anions in the solution.
For example, if 1 mole of NaCl is dissolved in water, there is 1 Eq of Na+ ions and 1 Eq of Cl- ions in the solution. If one mole of
CaCl2 is dissolved in the water there is 2 Eq of Ca2+ ions and 2 Eq of Cl- ions in the solution. Similarly, if 1 mole of NaCl and 1
mole of AlCl3 are dissolved in the water there is 1 Eq of Na+ ions, 3 Eq of Al3+ ions, and 4 Eq of Cl- ions to balance cations in the
water.
Equality gives two conversion factors. For example, the equality 1 mol Ca2+ = 2 Eq Ca2+ gives the following two conversion
factors:
2+
1 mol C a
2+
2 Eq C a
and
2+
2 Ea C a
2+
1 mol C a
The first conversion factor converts the given amount in equivalents to moles, and the second converts the given amount in moles
to equivalents of the ion as explained in the following examples.
Example 5.3.1
a) Calculate mEq of Fe3+ in 0.0200 mol of Fe3+? b) if chloride ion is the only anion in the solution, how many mEq of Cl- are
present in the solution?
Solution
a) Given 0.0200 mol Fe3+. Desired mEq Fe3+.
3+
1 mol F e3+
that converts the given amount in moles to Eq of
3+
3+ 3+ 3+
, that converts the Eq of Fe3+ to
1000 mEq F e
Fe . Another equality: 1 Eq Fe = 1000 mEq Fe gives the conversion factor 1 Eq F e
3+
mEq of Fe3+.
Calculations:
3+
3 Eq F e 3+
1000 mEq F e
3+ 3+
0.0200 mol F e × × = 60 mEq F e
1 mol F e3+ 1 Eq F e3+
Example 5.3.2
An intervenous saline solution contains 145 mEq/L of Na+. How many moles of Na+ are in 0.500 L of the solution?
Solution
a) Given: 145 mEq Na+ /L solution and 0.500 L solution, Desired: ? mol Na+
Conversion factors needed: The concentration in mEq/L is the first factor for L to mEq conversion, the equality 1 mEq Na+ = 1
mmol Na+ givens the second conversion needed for mEq to mmol conversion, and the equality 1000 mmol Na+ = 1 mol Na+
5.3.3 https://chem.libretexts.org/@go/page/372210
given the third conversion factor needed needed for mEq to mol conversion. The three conversion factors are applied one after
the other in a single row in the following calculation:
+
145 mEq Na 1 mmol Na
+
1 mol Na
+
0.500 L solution × × × = 0.0725 mol Na
+ +
1 L solution 1 mEq Na 1000 mmol Na
Note how the units of the numerator in one fraction are canceled by the units of the denumerator of the following fraction leaving
only the desired units uncancelled that become the units of the answer number.
The overall concentration of electrolytes in intravenous fluids given to patients is about the same as of electrolytes in the body
fluids. For example, Ringer’s lactate solution contains about: 130 mEq/L Na+, 4 mEq/L K+, 3 mEq/L Ca2+, 109 mEq/L Cl- , and 28
mEq/L lactate-. Note that the overall +ve charge (130+4+3= 137 mEq/L) is equal to the overall –ve charge (109+28 = 137 mEq/L).
5.3: Electrolytes is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
5.3.4 https://chem.libretexts.org/@go/page/372210
5.4: Concentration of solutions
The concentration of a solution tells the amount of solute dissolved in a given amount of solution.
Figure 5.4.1 : Making an aqueous solution: 1) measure the mass of the solute on an analytical balance, 2) pour the solute into a
volumetric flask of know volume, 3) add distilled water (or solvent) to more than half fill the flask, 4) stopper the flask and shake it
to dissolve the solute, 5) add more water to fill the flask to the calibration mark –measure from the lowest point of the water
meniscus, 6) stopper the flask and shake again to make a homogeneous solution. Source: modified from
https://www.hiclipart.com/free-trans...cvcrq/download
For example, 5% means 5:100, where 5 is the part, and 100 is the total. The percentage is calculated as a hundred times of part by
total, i.e.,
part
Percentage (%) = × 100.
Total
Example 5.4.1
A 50.0 g NaCl is dissolved in water to make a 500 g solution. What is the percentage of NaCl in the solution?
Solution
Given part = 50.0 g NaCl, and total = 500 g solution. Desired: %NaCl in the solution?
part
Formula: Percentage (%) = Total
× 100.
50.0cancel g
Calculations: Percentage (%) = 500cancel g
× 100 = 10.0 %NaCl
The units cancel in the fraction calculation part, and a % sign is added to the answer to tell that it is a fraction out of a hundred.
5.4.1 https://chem.libretexts.org/@go/page/372211
Note that the total is solute and solvent added together, i.e., solution.
Example 5.4.2
What is the mass % of NaOH in a solution prepared by dissolving 10.0 g NaOH in 100 g water?
Solution
Given solute = 10.0 g, and solvent = 100 g, Desired: Mass% NaOH?
mass of solute (g)
Formula: Mass (%) = mass of solute (g)+ mass of solvent (g)
× 100
10.0 gNaOH
Calculations: Mass (%) = (10.0 g+100 g) solution
× 100 = 9.09%NaOH
given g solute 100g solution
Note that the mass% concentration and its reciprocal are two conversion factor: 100g solution
and given g solute
Example 5.4.3
Neosporin antibiotic is a 3.5% m/m neomycin solution. How many grams of neomycin are in 50 g of ointment?
Solution
3.5 g neomycin
Given: 3.5% neomycin = 100 g solution
, and Solution amount = 50 g, Desired: ? g neomycin
Calculations:
3.5 g neomycin
50 g solution × = 1.8 g neomycin.
100 g solution
Fig. 5.4.2 shows the volume percent concentration ranges of different classes of fragrances.
Figure 5.4.2 : This info-graphic explains the differences between Parfums, Perfumes, Eau de Perfume, Eau de Parfum, Eau de
Toilette, and Cologne. It also gives an approximate idea of how the concentration of Parfum oils vary as we move to lighter scents.
Source: Nicole Smith / CC BY-SA (https://creativecommons.org/licenses/by-sa/4.0)
5.4.2 https://chem.libretexts.org/@go/page/372211
Example 5.4.4
What is the volume % of rose extract in a solution prepared by dissolving 14.0 mL rose extract in a solvent to make 200 mL of
solution?
Solution
Given: Solute = 14.0 g, and solution = 200 g, Desired: Volume% rose solution?
Formula:
volume of solute (mL)
Volume (%) = × 100
volume of solution (mL)
Calculations:
14 mL rose extract
Volume (%) = × 100 = 7.0% rose solution
200 mL solution
Example 5.4.5
What is the volume of bromine (Br2) in 250 mL of 4.8% v/v of Br2 solution in carbon tetrachloride?
Solution
Given: Concentration 4.8% v/v bromine = 4.8 mL bromine
100 mL solution
, volume of solution = 250 mL, Desired: Volume of solute, i.e., ? mL
bromine.
Calculations:
4.8 mL bromine
250 mL solution × = 12 mL bromine.
100 mL solution.
Example 5.4.6
What is the mass/volume % of glucose solution prepared by dissolving 50 g glucose in enough water to make 1000 mL of
solution?
Solution
Given: Solute = 50.0 g, and Solution = 1000 mL, Desired: Mass/volume % glucose solution?
Formula:
mass of solute (g)
Mass/volume (%) = × 100
volume of solution (mL)
Calculations:
5.4.3 https://chem.libretexts.org/@go/page/372211
Mass 50 g glucose m
(%) = × 100 = 5.0% glucose solution by .
volume 1000 mL solution v
Example 5.4.7
How many grams of clindamycin antibiotics are in a 45 mL capsule of the 1.0% (m/v) clindamycin?
Solution
1.0 g clindamycin
Given: % m/v concentration: 1.08 % m/v clindamycin =
100 mL solution
, and volume of solution = 45 mL, Desired: ? g
clindamycin?
Calculations:
1.0 g clindamycin
45 mL solution × = 4.5 g clindamycin.
100 mL solution
Parts per million (ppm) and parts per billion (ppb) concentration
Parts per million (ppm) is a number or ratio expressed as a fraction of a million (106).
1,000,000
or 2:1,000,000, where 2 is the part, and 1,000,000 is the total. The concentration in ppm is
calculated as a million times of part by total, i.e.:
part
6
Concentration in ppm = × 10
Total
Parts per billion (ppb) is a number or ratio expressed as a fraction of billion (109).
That is:
part
9
Concentration in ppb = × 10
Total
Like percentage concentration, the ppm and ppb can be mass/mass (m/m), volume/volume (v/v) or mass/volume (m/v).
Example 5.4.8
EPA’s action limit for copper is 1.3 mg/L in drinking water. What is this limit in ppm of copper m/v in the drinking water?
Solution
Given: 1.3 mg copper in 1L solution, Desired: ? ppm m/v of Copper in water
solute( g)
Formula: Concentration in ppm = solution( mL)
6
× 10
Calculations: First, convert the given units of mass and volume into the corresponding units that the formula takes, then plug
the values in the formula and calculate.
1 g
Solute = 1.3 mg × = 0.0013 g
1000 mg
1000 mL
Solution = 1 L × = 1000 mL
1 L
0.0013 g
6
Concentration in ppm = × 10 = 1.3 ppm copper v/m
1000 mL
5.4.4 https://chem.libretexts.org/@go/page/372211
Example 5.4.9
EPA’s action limit for lead is 0.015 mg/L in drinking water. What is this limit in ppb of lead m/v in the drinking water?
Solution
Given: 0.015 mg in 1L solution, Desired: ? ppb m/v of Lead in water
solute(g)
Formula: Concentration in ppb = solution(mL)
9
× 10
Calculations: First, convert the given units of mass and volume into the corresponding units that the formula takes, then plug
the values in the formula and calculate.
1 g
Solute = 0.015 mg × = 0.000015 g
1000 mg
1000 mL
Solution = 1 L × = 1000 mL
1 L
0.000015 g
9
Concentration in ppb = × 10 = 15 ppb lead v/m
1000 mL
Molarity
Molarity (M) expresses the moles of solute in a liter of solution.
The most common solution concentration unit used in chemistry is molarity (M):
n (moles of solute )
Molarity (M ) =
V (Litters of solution )
Example 5.4.10
What is the molarity (M) of a solution prepared by dissolving 50.0 g NaOH in enough water to make 250 mL solution?
Solution
Given: Solute = 50.0 g NaOH, Volume of solution = 250 mL, Desired: ? M NaOH solution?
n (moles of solute )
Formula: Molarity (M ) = V (Litters of solution )
Calculations: First, convert the given units of mass and volume into the corresponding units that the formula takes, then plug
the values in the formula and calculate.
1 mol NaOH
Solute = 50.0 g NaOH × = 1.25 mol NaOH
40.00 g NaOH
1 L
Solution = 250 mL × = 0.250 L
1000 mL
1.25 mol NaOH
Molarity (M ) = = 5.00 M NaOH
0.250 L solution
Example 5.4.11
How many litters of 0.211 M HCl solution are needed to provide 0.400 mol of HCl?
Solution
5.4.5 https://chem.libretexts.org/@go/page/372211
Given: amount of solute = 0.400 mol HCl, Concentration of solute = 0.211M = 0.211 mol HCl
1L solution
Calculations:
1 L soluton
Liters of solution needed = 0.400 mol HCl × = 1.90 L solution
0.211 mol HCl
Example 5.4.12
1 L solution
, Desired: ? moles of
NaOH?
Calculations:
1.12 mol NaOH
Moles of NaOH in the solution = 2.50 L solution × = 2.80 mol NaOH
1 L solution
Dilution of solutions
Dilution of a solution is the addition of a solvent to decrease the solute concentration of the solute in the solution.
The product of concentration (C) and volume (V) is the amount of solute, i.e.,
amount of solute
Amount of solute = C in × V in volume of solution.
volume of solution
The amount of solute does not change by adding solvent. Therefore, the product of concentration and volume, i.e., CV, which is the
amount of solute, is a constant, i.e.,
C1 V1 = C2 V2 = amount of solute
Figure 5.4.3 : This image illustrates that when more solvent is mixed with the solution, the amount of solute remains the same, but
its concentration decreases. Source: Theislikerice / CC BY-SA (https://creativecommons.org/licenses/by-sa/4.0)
Fig. 5.4.3 shows that if the initial concentration is C1, the initial volume is V1, and after dilution, the final concentration is C2, the
final volume is V2, then C1V1 = C2V2 = amount of solute that is constant. If three of the four variables in this equation are known,
the missing one can be calculated, as explained in the following example.
Caution
Keep in mind that the concentrations and volumes should be in the same units on both sides of the equation: C1V1 = C2V2. If
they are not in the same units, convert them to the same units before plunging them in the formula.
5.4.6 https://chem.libretexts.org/@go/page/372211
Example 5.4.13
how much volume of 11.3 M HCl is needed to prepare 250 mL of 2.00 M HCl?
Solution
Given: C1 = 11.3 M HCl, C2 = 2.00 M HCl, V2 = 250 mL solution, Desired V1 = ?
C2 V2
Formula: C1V1 = C2V2, rearrange it to isolate the desired parameter: V 1 =
C1
Calculations:
C2 V2 2.00 M HCl × 250 mL solution
V1 = = = 44.2 mL solution
C1 11.3 M HCl
Example 5.4.14
what is the molarity of the NaOH solution prepared by diluting 100 mL of 0.521 M NaOH solution to 500 mL?
Solution
Given C1 = 0.521 M NaOH, V1 = 100 mL solution, V2 = 500 mL solution, Desired: Concentration of the final solution C2 = ?
M NaOH
C1 V1
Formula: C1V1 = C2V2, rearrange to isolate the desired parameter: C 2 =
V2
Calculations:
C1 V1 0.521 M NaOH × 100 mL NaOH
C2 = = = 0.104 M NaOH
V2 500 mL NaOH
Example 5.4.15
Dopamine is administered intravenously to a patient to increase blood pressure. How many milliliters (mL) of a 4.0% (m/v)
dopamine solution is needed to prepare 250 mL of a 0.030% m/v) solution?
Solution
Given: C1 = 4.0% (m/v), C2 = 0.030% (m/v), V2 = 250 mL solution, Desired V1= ?
Formula: C1V1 = C2V2, rearrange to isolate the desired parameter: V 1 =
C2 V2
C1
Calculations:
C2 V2 0.030% × 250 mL solution
V1 = = = 5.0 mL of solution
C1 4.0%
Logarithmic dilution
Take a unit volume of a given solution and add enough solvent to increase the volume of the solution 10 times for a
logarithmic diluiton.
A logarithmic dilution is ten times dilution, i.e., proven by the following formula:
C1 V1 1 mL × C1
C2 = = = 0.1 × C1
V2 10 mL
Repeating the above step with the diluted solution results in 10x10 = 100-time dilution, and repeating third-time results in
10x10x10 = 1000-time dilution. This dilution of 10 times in each step is called logarithmic dilution. Fig. 5.4.4 shows that five steps
of logarithmic dilution on a 10% initial solution results in a concentration of 10 ppm in the final solution.
5.4.7 https://chem.libretexts.org/@go/page/372211
Figure 5.4.4 : Logarithmic dilution: five steps of logarithmic dilution on a 10% initial solution results in a concentration of 10 ppm
of the final solution. Source: Grasso Luigi / CC BY-SA (https://creativecommons.org/licenses/by-sa/4.0)
5.4: Concentration of solutions is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
5.4.8 https://chem.libretexts.org/@go/page/372211
5.5: Osmosis
Semipermeable membranes surround living cells and organelles in the cells. The semipermeable membranes allow water and small
molecules but do not allow the passage of large molecules and ions.
Osmosis is the passage of water and small molecules across a semipermeable membrane with a net flow from a less
concentrated solution to a more concentrated solution.
Osmosis helps in the absorption, retention, and flow of water, nutrients, and other molecules required in the biological systems. Fig.
5.5.1 illustrates the process of osmosis.
Figure 5.5.1 : Illustration of osmosis. Osmosis is the passage of water and small molecules through a semipermeable membrane
with a net flow from low solute concentration to high solute concentration. The larger molecules and ions cannot pass the
semipermeable membrane.
Osmotic pressure
Osmosis causes net water flow from the less concentrated to the more concentrated solution across the semipermeable membrane.
Consequently, the water level rises in the less concentrated solution compartment, as illustrated in Fig. 5.5.2. The difference in the
height of water increases and applies pressure, pumping water back to the more concentrated side until the flow of water is equal
on the two sides.
Figure 5.5.2 : Osmotic pressure is the pressure, due to the difference in the height of water in this example, that prevents the excess
flow of water from less concentrated to more concentrated solution side across the semipermeable membrane that separates the two
solutions. Source: WYassineMrabetTalk✉ This W3C-unspecified vector image was created with Inkscape. / CC BY-SA
(https://creativecommons.org/licenses/by-sa/3.0)
5.5.1 https://chem.libretexts.org/@go/page/372212
The pressure that prevents additional water flow to the more concentrated solution side of the semipermeable membrane is
called osmotic pressure.
The osmotic pressure is proportional to the overall concentration of the solute particles. For example, 0.1 molar NaCl has osmotic
pressure about twice that of 0.1 molar glucose because each mole of glucose adds one mole of solute particles, while each mole of
NaCl produces two moles of particles, i.e., one-mole Na+ and one-mole Cl- ions in the solution.
Reverse Osmosis
Reverse osmosis is the net flow of water to the less concentrated or pure water across the semipermeable membrane by
applying external pressure more than the osmotic pressure on the more concentrated solution side.
Reverse osmosis is the net flow of water to the less concentrated or pure water across the semipermeable membrane by applying
external pressure more than the osmotic pressure. Note that solvent flow in reverse osmosis driven by external pressure is the
opposite of regular osmosis. Reverse osmosis is used to produce drinking water from seawater sources, as illustrated in Fig. 5.5.3.
Figure 5.5.3 : Illustration of reverse osmosis –used to produce drinking water from a seawater source. Source: Colby Fisher / CC
BY-SA (https://creativecommons.org/licenses/by-sa/3.0)
Solutions with the same solute particle concentration and osmotic pressure are called isotonic. If the two solutions across a
semipermeable membrane do not have the same solute particle concentration, the solution with higher solute particle
concentration and higher osmotic pressure is hypertonic, and the other has lower solute particle concentration and lower
osmotic pressure is hypotonic.
Remember that hyper- means more and hypo-means less, concerning solute particle concentration in the case of osmosis.
Figure 5.5.4 : A hypertonic solution has a solute concentration higher than another solution. An isotonic solution has a solute
concentration equal to another solution. A hypotonic solution has a solute concentration lower than another solution. Source:
OpenStax / CC BY (https://creativecommons.org/licenses/by/3.0)
Cells placed in an external solution may retain their size, shrink, or swell depending on the relative osmotic pressure of fluid inside
and outside of the cell, as illustrated in Fig. 5.5.4. For example, red blood cells placed in an isotonic solution retain their size
because the flow of water into and out of the cell is the same.
5.5.2 https://chem.libretexts.org/@go/page/372212
Typical isotonic solutions are 0.9% m/v NaCl solution in water or 5% m/v glucose solution in water.
Red blood cells placed in a hypertonic solution shrink in size due to more flow of water out than into the cell –a process
called crenation.
Red blood cells placed in a hypotonic solution swell and burst due to more water flow into than out of the cells –a process
called hemolysis.
A similar situation happens in plant cells that are placed in different environments concerning osmotic pressure, as illustrated In
Fig. 5.5.5.
Figure 5.5.5 : Plant cells under different environments concerning osmotic pressure. source: LadyofHats / Public domain
Dialysis
Dialysis separates colloids from water, dissolved ions, and molecules of small dimention
Dialysis is similar to osmosis with the difference that in dialysis water, small molecules and ions can pass through a dialyzing
membrane leaving behind collide particles like proteins and starch molecules, as illustrated in Fig. 5.5.6.
Figure 5.5.6 : Dialysis separates colloidal particles like protein and starch from water, small molecules, and ions. Source:
Potcherboy at English Wikipedia / CC BY (https://creativecommons.org/licenses/by/3.0)
5.5.3 https://chem.libretexts.org/@go/page/372212
Figure 5.5.7 : Schematics of dialysis for blood cleaning through kidney and nephron. 1: Renal cortex. 2: Medulla. 3: Renal artery. 4:
Renal vein. 5: Ureter. 6: Nephrons. 7: Afferent arteriole. 8: Glomerulus. 9: Bowman’s capsule. 10: Tubuli and loop of Henle. 11:
Efferent arteriole. 12: Peritubular capillaries. Source: File:Physiology_of_Nephron.svg:
Madhero88File:KidneyStructures_PioM.svg: Piotr Michał Jaworski; PioM EN DE PLderivative work: Daniel Sachse (Antares42) /
CC BY-SA (https://creativecommons.org/licenses/by-sa/3.0)
Hemodialysis
Hemodialysis is used to extract urea and other waste products from the blood when a person's kidney fails to remove them, as
illustrated in Fig. 5.5.8. A hemodialysis system is a kind of artificial kidney in which the blood flows through long cellophane tubes
placed in an isotonic solution containing NaCl, KCl, NaHCO3, and glucose. Cellophane is a dialyzing membrane that does not let
proteins, other large molecules, and blood pass through it, but urea excretes.
Figure 5.5.8 : Simplified hemodialysis circuit. Source: GYassineMrabetTalk✉ This W3C-unspecified vector image was created
with Inkscape. / CC BY (https://creativecommons.org/licenses/by/3.0)
5.5: Osmosis is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
5.5.4 https://chem.libretexts.org/@go/page/372212
CHAPTER OVERVIEW
6: Acids and bases
6.1: What is an acid and a base?
6.2: Brønsted–Lowry acids and bases
6.3: Strength of acids and bases
6.4: Acid-base equilibrium
6.5: Dissociation of water
6.6: The pH
6.7: Acid-base reactions
6.8: pH Buffers
6: Acids and bases is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
1
6.1: What is an acid and a base?
General properties of acids and bases
We commonly encounter acids and bases in our foods –some foods are acidic, and others are basic (alkaline), as illustrated in Fig.
6.1.1.
Figure 6.1.1 : Acidic and alkaline foods along with a pH scale. Source: Illustration by Cam Howard, 2016 / CC0
The general properties of acids and bases are the following.
1. Acids taste sour, e.g., citrus fruits taste source because of citrus acid and ascorbic acid, i.e., vitamin C, in them. Basic
(alkaline) substances, on the other hand, taste bitter.
2. Basic (alkaline) substances feel soupy, while acidic substances may sting.
3. The acids turn blue litmus paper to read but do not change the color of red litmus paper. Bases turn red litmus paper blue
but do not change the color of blue litmus paper, as illustrated in Fig. 6.1.2.
4. Phenolphthalein indicator turns colorless in acid and turns pink in basic solution, as illustrated in Fig. 6.1.3.
5. Acids and bases neutralize each other. Hydrochloric acid is found in the stomach that helps digestion. Excess hydrochloric
acid may cause acid burns—antacids like milk of magnesia are bases that help by neutralizing excess acid in the stomach.
Figure 6.1.2 : Demonstration of acids turn red litmus paper blue and bases turn blue litmus paper red - solution listed at the top of
the image was spotted on the red litmus paper (top) or blue litmus paper (bottom).
Figure 6.1.3 : Colours of phenolphthalein in acid(left) and base (right) solutions. Source: User:Siegert / Public domain.
An acid is a substance that forms hydrogen ions H+ when dissolved in water, and
A base is a substance that forms hydroxide ions OH- when dissolved in water.
6.1.1 https://chem.libretexts.org/@go/page/371798
or example, hydrochloric acid is an acid because it forms H+ when it dissolves in water.
Water
+ −
HCl(g) ⟶ H (aq) + Cl (aq)
Note that hydrogen ion H+ does not exist in reality. It bonds with water molecules and exists as hydronium ion H3O+(aq).
+ +
H (aq) + H2 O → H3 O (aq)
6.1.2 https://chem.libretexts.org/@go/page/371798
Acid formula Acid name Anion Anion name
Table 2 lists the names and formulas of some of the common Arrhenius bases.
The Arrhenius bases are ionic compounds of metal and hydroxide ion, and their name starts with the name of the metal element
followed by the name of the anion, i.e., hydroxide. For example, NaOH is sodium hydroxide.
6.1: What is an acid and a base? is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
6.1.3 https://chem.libretexts.org/@go/page/371798
6.2: Brønsted–Lowry acids and bases
Some bases do not have hydroxide ions in their formula, yet they act as bases and neutralize acids. For example, ammonia (NH3)
and calcium carbonate (CaCO3) do not contain hydroxide ions, but they neutralize acids. Further, Arrhenius's definition limits the
acid-base reactions in the water medium. The acid-base reactions can take place in other mediums also, e.g., HCl –an acid, and
NH3 –a base can react with and neutralize each other in the gas phase also. The Brønsted–Lowry bordered the definition of acids
and bases by including the bases mentioned above and also by including acid-base reactions in a non-aqueous medium.
Water is a base in the above reaction because it accepts a proton from the acid. In a reaction between HCl and NH3:
HCl is an acid because it donates its proton to NH3, and the NH3 is a base because it accepts a proton, as shown in Fig. 6.2.1.
Figure 6.2.1 : Proton transfer from HCl to NH3 is an example of an acid-base reaction. Hydrochloric acid evaporates from liquid
HCl in a beaker reacting with ammonia fumes coming from the test tube to produce ammonium chloride (white smoke). Source:
Walkerma, Public domain.
Brønsted–Lowry's acids have ionizable protons that they donate to bases. Therefore, Brønsted–Lowry's acid is generally written as
HA, where H+ is the donatable proton, and A- is the anion of the acid. Examples of acids are HCl, H2SO4, HNO3, and CH3COOH.
Note that acetic acid has only one acidic proton that is attached to the O atom in the carboxylic acid group (–COOH). The rest of
the protons attached to carbon atoms are not acidic. All organic acids have a carboxylic acid group (–COOH). Brønsted–Lowry's
acid may have net +ve charge, no charge, or net –ve charge on it. For example, H3O+, HCl, and HSO4- are all acids because they
can donate a proton to a base.
6.2.1 https://chem.libretexts.org/@go/page/371799
The base accepts proton by making a bond with it. The bond is a pair of bonded electrons. Since the proton is a hydrogen atom
without an electron, both electrons in the bond come from the base. The base must have a lone pair of electrons on it. The base is
usually represented as to emphasize a lone pair of electrons on it that is shown as a pair of dots. For example, ammonia, water,
and hydroxide ion ( , , and ) are Brønsted–Lowry's bases, because each of these has an atom with lone pair or lone
pairs of electrons on them.
The conjugate acid-base pair is related to the loss and gain of H+. For example, HF/F- is a conjugate acid-base pair, and
H3O+/H2O is also a conjugate acid-base pair.
In other words, remove the acidic proton from an acid to get its conjugate base and add a proton to a base to get its conjugate acid.
Figure 6.2.2 : HF –an acid that loses a proton to form its conjugate base F-, and water -acts as a base by accepting a proton to form
its conjugate acid H3O+.
Another example is ammonia NH3 which dissolves in water and accepts a proton to form its conjugate acid NH4+, as shown in Fig.
6.2.3. Water H2O acts as an acid by donating a proton and forming its conjugate base OH-. The two conjugate acid-base pairs in
this reaction are NH4+ /NH3 and H2O/OH- that are related by loss and gain or an H+.
Figure 6.2.3 : H2O –acts as an acid that loses a proton to form its conjugate base OH-, and NH3 –a base accepts a proton to form its
conjugate acid NH4+.
6.2.2 https://chem.libretexts.org/@go/page/371799
Example 6.2.1
Solution
1. Identify the substance that has donated a proton in the reactants –it is an acid.
2. Remove a proton from the acid to form its conjugate base: H3PO4/H2PO4-.
3. Identify the substance that has accepted a proton in the reactants –it is a base.
4. Add a proton to the base to form its conjugate acid: NH3/NH4+.
Note
Note that loss of a proton from an acid forms its conjugated base with the charge decreased by one, e.g., H3PO4/H2PO4-,
HSO4-/SO42-, and NH4+/NH3. Similarly, the gain of a proton by a base forms its conjugate acid with the charge increased by
one, e.g., HPO42-/H2PO4-, HCO3-/H2CO3, and NH3/NH4+.
Amphoteric substances
Water acts as a base in some reactions, e.g., with HF, and as an acid in some reactions, e.g., with NH3.
Substances like water that can act as an acid and also as a base are called amphoteric substances.
Other examples of amphoteric substances include HSO4-, HCO3-, and NH3, as illustrated in Fig. 6.2.4.
Figure 6.2.4 : Examples of amphoteric substances that can act as an acid in some reactions and also as a base in other reactions.
6.2: Brønsted–Lowry acids and bases is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif
Malik.
6.2.3 https://chem.libretexts.org/@go/page/371799
6.3: Strength of acids and bases
The strength of acid HA is the extent to which the acid dissociates into H+ and A- ions, as illustrated in Fig. 6.3.1.
Figure 6.3.1 : Image of strong acid mostly dissociating (left) and a weak acid partially dissociating into ions in water (right).
Source: Cwszot / CC0
Strong acids
Strong acids, like HCl, almost 100% dissociate into ions when they dissolve in water.
+ −
HCl(g) + H2 O(l) → H3 O (aq) + Cl (aq)
One arrow is used to indicate that the reaction is nearly 100% complete.
Strong acids include HClO4, H2SO4, HI, HBr, HCl, and HNO3
Weak acids
Weak acids dissolve in water but partially dissociate into ions.
For example, acetic acid (CH3COOH) is a weak acid, 1 M acetic acid dissolves in water, but only 0.4% of the dissolved molecules
dissociate into ions, the remaining 99.6% remain undissociated, as illustrated in Fig. 6.3.2. and equation of the dissociation
equilibrium below.
+ −
CH3 COOH(aq) + H2 O(I) ⟵⃗ H O (aq) + CH3 COO (aq)
3
Two arrows pointing in opposite directions are used for the dissociation of weak acids to indicate that the reaction is an
equilibrium, i.e., two ways.
Often the arrows are not equal in size -the longer arrow points to acid-base pair that is weaker and present in a larger concentration
at equilibrium than their conjugate pair.
6.3.1 https://chem.libretexts.org/@go/page/371800
Figure 6.3.2 : After dissolution in water. HCl –the strong acid is 100% dissociated into H+ and Cl- ions leaving no dissolved HCl
molecules in water, but acetic acid (HAc) –the weak acid has a high concentration of HAc molecules and low concentration of H+
and Ac- ions.
Strong bases
Strong bases almost %100 dissociate into ions when dissolved in water. For example, NaOH is a strong base, and it dissociates
almost 100% into ions in water.
Strong bases almost %100 dissociate into ions when dissolved in water. For example, NaOH is a strong base, and it dissociates
almost 100% into ions in water.
Water
+ −
NaOH(s) ⟶ Na (aq) + OH (aq)
One arrow is used for the dissolution of strong bases to indicate that the reaction is almost complete.
Strong bases include hydroxides of alkali metals, i.e., LiOH, NaOH, KOH, RbOH, CsOH, and hydroxides of heavy alkaline
earth metals, i.e., Ca(OH)2, Sr(OH)2, and Ba(OH)2.
The last three, i.e., the hydroxides of heavy alkaline earth metals, have low solubility in water, but the dissolved fraction exists as
ions.
Weak bases
Weak bases partially dissociate into ions when dissolved in water.
For example, ammonia is a weak base –only 0.42% of the dissolved ammonia molecules dissociate into ammonium ions and
hydroxide ions in water from a 1 M solution of ammonia.
−
NH3 (aq) + H2 O(I) ⟵⃗ NH + (aq) + OH (aq)
4
Weak bases in household use include ammonia (NH3) in window cleaners, NaClO in bleach, Na2CO3 and Na3PO4 in laundry
detergent, NaHCO3 in tooth past, Na2CO3 in baking powder, CaCO3 for use in lawns, Mg(OH)2 and Al(OH)3 in antacids and
laxatives.
The weak bases mentioned above are all ionic compounds except ammonia. Ionic compounds are strong electrolytes, i.e., they
dissociate into ions almost 100% upon dissolution in water. It appears to contradict the fact that these ionic compounds are weak
bases. It does not actually contradict, because the base properties do not refer to these ionic compounds, the base properties refer to
the reactions of their polyatomic anions, i.e., ClO-, CO32-, and PO43- with water, as shown in the reactions below:
− −
ClO + H2 O ⟵⃗ HClO + OH
2− −
CO + 2 H2 O ⟵⃗ H CO + 2 OH , and
3 2 3
3− −
PO4 + 3 H2 O ⟵⃗ H PO + 3 OH
3 4
6.3.2 https://chem.libretexts.org/@go/page/371800
The above reactions are equilibrium reactions that are more favored in the revers than the forward direction, producing a small
number of OH- ions compared to the anion on the reactant sides. The last two examples, i.e., Mg(OH)2 and Al(OH)2 are classified
as weak bases because they are considered insoluble in water. The solubility of Mg(OH)2 is 0.00064 g/100 mL (25 °C), and the
solubility of Al(OH)3 is 0.0001 g/100 mL, which are in the range of insoluble ionic compounds.
The solubility and the strength of acids and bases are two different things. A strong base may be less soluble, and a weak base
may be more soluble or vice versa, but a dissolved strong base exists as ions only, and a dissolved weak base exists both as
molecules and ions.
The conjugate bases of strong acids have negligible base strength, and the conjugate acids of strong basses have negligible acid
strength. Fig. 6.3.3. illustrates the relative strengths of some acids and their conjugated bases.
For example, a dissociation reaction between HCl and H2O is almost 100% complete because HCl is a stronger acid than H3O+ and
H2O is a stronger base than Cl-:
+ −
HCl + H2 O → H3 O + Cl
6.3.3 https://chem.libretexts.org/@go/page/371800
The dissolution of acetic acid (CH3COOH) and ammonia (NH3) are equilibrium reactions because all the acids, bases, and their
conjugates are in the weak acids or weak bases category. However, acetic acid and water dominate over their conjugates H3O+ and
CH3COO- by 99.6:0.4 ratio (in 1 M acetic acid solution) because the conjugate acid H3O+ is a stronger acid than CH3COOH, and
conjugate base CH3COO- is a stronger base than H2O.
+ −
CH3 COOH(aq) + H2 O(l) ⟵⃗ H O (aq) + CH3 COO (aq)
3
The longer arrow, in the unbalanced equilibrium arrows, points to the acid-base pair in the reaction that exists in a higher
concentration relative to their conjugates.
Similarly, ammonia (NH3) and water (H2O) dominate over their conjugates NH4+ and OH- by ~99.6:0.4 ratio (1M ammonia
solution) because the conjugate acid NH4+ is a stronger acid than H2O and conjugate base OH- is a stronger base than NH3.
+ −
NH3 (aq) + H2 O(I)) ⟵⃗ NH (aq) + OH (aq)
4
6.3: Strength of acids and bases is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
6.3.4 https://chem.libretexts.org/@go/page/371800
6.4: Acid-base equilibrium
Most of the acid-base reactions are equilibrium reactions, i.e., the reactants form the products, and the products react to re-form the
reactants. The reaction is shown with double arrows to indicate that both the forward and the reverse reactions happen
simultaneously.
At the beginning of the equilibrium when the rate of the reverse reaction is slower than the rate of forward reaction is the
pre-equilibrium phase.
The equilibrium phaes starts at the point when the rate of reverse reaction becomes equal to the rate of the forward reaction
Figure 6.4.1 : In pre-equilibrium, reactants react and their concentration, and, consequently, the rate of forward reaction decreases
over time. The products build-up and the reverse reaction increases over time. When rate of forward reaction becomes equal to the
rate of reverse reaction the equilibrium is reached and the concentration of reactants and products do not change.
At the point when the rate of reverse reaction becomes equal to the rate of the forward reaction, equilibrium has reached. The
concentration of the reactants and products do not change at equilibrium because they are consumed and re-formed at the same rate
–it is a dynamic equilibrium.
The equilibrium reactions are not limited to acid-base reactions; they are common in all types of chemical reactions. Fig. 6.4.2
illustrates the concentrations and rates of reactions changes in pre-equilibrium and equilibrium phases with the help of an actual
chemical equilibrium reaction between a decomposition reaction of a colorless gas N2O4 and its reverse reaction, i.e., a
combination reaction of brown color gas NO2.
6.4.1 https://chem.libretexts.org/@go/page/371801
Figure 6.4.2 : Top row: a sealed tube containing colorless gas N2O4 darkens as it decomposes to yield brown gas NO2 by the
reaction: N O ⇄ 2NO . The middle row: shows the models of the molecules at the timeline marked by the dotted lines in the
2 4 2
graphs. Bottom row: shows concentration vs time (left graph); rate of forward reaction by blue curve and rate of reverse reaction by
red curve (right graph). Source: OpenStax / CC BY (https://creativecommons.org/licenses/by/4.0)
Le Chatelier's principle
If a chemical equilibrium is disturbed, the rates of forward and reverse reactions change to relieve the stress and re-establish
the equilibrium.
The stresses can be changes in the concentration, pressure, or temperature, as explained in the following section.
Fig. 6.4.3 illustrates the effects with the help of water level in two tanks connected through a conduit in the left tank representing
reactants and in the right tank representing products.
6.4.2 https://chem.libretexts.org/@go/page/371801
Figure 6.4.3 : Illustration of Le Chatelier's principle using two water tanks connected through a conduit; 1) at equilibrium the rate
of forward and reverse reactions are equal (the topmost), 2) when any reactant is increased the forward reaction rate increase
(middle left), 3) when any product is increased the reverse reaction rate increases (middle right), 4) when any reactant is decreased
the forward reaction rate decreases (bottom left), and 5) when any product is decreased the reverse reaction decreases (bottom
right)
Fig 6.4.4 demonstrates the effect of concentration change on the chemical equilibrium between brown dichromate ions and yellow
chromate ions:
the equilibrium to left making more of brown Cr2O72- shown in the middle tube. Removal of H+ by adding a base NaOH shift the
reaction to the right making more of yellew CrO42- shown in right most tube. Source: NCSSMDistanceEd / CC BY
(https://creativecommons.org/licenses/by/3.0)
2− 2− +
Cr2 O (aq, brown ) + H2 O(I) ⇄ 2 CrO (aq, yellow ) + 2 H (aq)
7 4
The addition of acid to the test tube on the left increases H+ in the system, increasing reverse reaction, which can be observed by
the increased brown color in the middle test tube. Then, the addition of a base removes H+ from the system by an acid-base
reaction: H + OH → H O . The decrease in H+ shifts the equilibrium to the product side, which can be observed by the
+ −
2
6.4.3 https://chem.libretexts.org/@go/page/371801
Hb(aq) + O2 ( g) ⇄ HbO 2 (aq)
The concentration of oxygen is higher in the lunges that shift the equilibrium to the product side, binding more oxygen with the
hemoglobin. When the blood arrives in tissues, the concentration of oxygen is lower in tissues, causing the equilibrium to shift to
the reactant side, releasing the oxygen.
The oxygen concentration decreases as the altitude increases. The mountain climbers may experience hypoxia, i.e., inadequate
supply of oxygen to the body because the lower level of oxygen at high altitude may shift the equilibrium to the right, resulting in
less binding of oxygen with the hemoglobin in the lungs. The body reacts by producing more hemoglobin, but it takes about 10
days for the body to re-adjust the hemoglobin level in the blood. Peoples living at higher altitudes usually have a higher level of
hemoglobin in their blood for the reasons described above.
Figure 6.4.5 : Illustration of increase in pressure causes decrease in the volume and increase in the concentration of gas molecules.
Source: Olivier Cleynen / CC0
If a chemical equilibrium involves gases, an increase in pressure has the same effect as increases in the gaseous reactant or
product—the equilibrium shifts in the direction where there are fewer moles of gases.
Fig. 6.4.6 demonstrated this effect for the equilibrium between colorless N2O4 gas and brown color NO2 gas:
The color becomes light in going from a syringe on the right to the syringe in the middle. It can be explained based on the fact that
the concentration of gases decreased upon an increase in volume as a result of a decrease in pressure. Then the color becomes
darker over time, as shown by the syringe on the right. The darker color indicates that the equilibrium has shifted to the product
side where there are more moles of gas, to relieve the stress.
Figure 6.4.6 : Decrease in pressure on the equilibrium: N2 O4 ( g, colorless ) ⇄ 2NO2 ( g, brown ), Middle syringe → syringe on
the right, increases the forwared reaction to release the stress. Source: NCSSMDistanceEd / CC BY
(https://creativecommons.org/licenses/by/3.0)
6.4.4 https://chem.libretexts.org/@go/page/371801
When a reversible reaction at equilibrium is disturbed by increasing temperature, the equilibrium shifts in the endothermic
direction of the reaction to remove the heat and vice versa.
In the equilibrium between colorless N2O4 and brown color NO2 gases, the reaction is endothermic in the forward direction:
An increase in temperature shifts the equilibrium in the endothermic direction to relieve the stress, as demonstrated in Fig. 6.4.7.
Figure 6.4.7 : Increase in temperature shifts the equilibrium in the endothermic direction producing more of the colored NO2 in the
following equilibrium: N O (g, colorless ) + Heat ⇄ 2NO (g, brown ). An overlay of the same 99.9% pure NO2/N2O4
2 4 2
sealed in an ampoule. From left to right -196 °C, 0 °C, 23 °C, 35 °C, 50 °C. Source: Eframgoldberg / CC BY-SA
(https://creativecommons.org/licenses/by-sa/3.0)
6.4: Acid-base equilibrium is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
6.4.5 https://chem.libretexts.org/@go/page/371801
6.5: Dissociation of water
How do water molecules dissociate?
Water is an amphoteric substance, which means water can accept a proton acting as a base, and it can also donate A proton acting
as an acid.
About one water molecule in half a billion dissociates into an OH- ion by losing a proton to another water molecule. The
molecule that receives a proton becomes H3O+.
The dissociation of water is an equilibrium reaction in which one water molecule donates its proton to another water molecule.
The water molecule that receives proton is acting as a base, and it converts to conjugate acid H3O+. The other water molecule that
donates a proton is acting as an acid, and it converts to conjugate base OH-. The arrows in the reaction show that the base uses one
of its lone pairs of electrons to make a bond with proton, and the previous bond pair of electrons turns into a third lone pair of
electrons on the oxygen atom of the base. The reaction is reversible, i.e., the conjugate acid (H3O+) and the conjugate base (OH-)
react to re-form the two water molecules.
The product of the molar concentration of H3O+ and OH- in water is a constant called water dissociation constant Kw equal
to 10-14 at 25 oC, i.e.:
+ − −7 −7 −14
Kw = [ H3 O ] [ OH ] = (10 ) (10 ) = 10 (6.5.1)
An acidic solution has an acid dissolved in water. When an acid dissolves in water it dissociates adding more H3O+. The [OH-]
must decrease to keep the Kw constant.
A solution that has [H3O+] more than 10-7, and [OH-] less than 10-7 is an acidic solution.
A basic solution has a base dissolved in water. When a base dissolves in water it dissociates adding more OH-. The [H3O+] must
decrease to keep the Kw constant.
A solution that has [H3O+] less than 10-7, and [OH-] more than 10-7 is a basic solution.
Therefore, if the molar concentration of hydronium ions [H3O+] is known, the molar concentration of hydroxide ions [OH-] can be
calculated using the following formula:
6.5.1 https://chem.libretexts.org/@go/page/371806
−14
Kw 10
+
[ H3 O ] = =
− −
[ OH ] [ OH ]
Similarly, if the molar concentration of hydroxide ions [OH-] is known, the molar concentration of hydronium ions [OH-] can be
calculated using the following formula:
−14
−
Kw 10
[ OH ] = =
+ +
[ H3 O ] [ H3 O ]
When a strong acid like HCl dissolves in water, it dissociates ~100% into ions. Therefore, the [H3O+] is equal to the molar
concentration of the acid. The amount H3O+ added by dissociation of water molecules is very small compared to that coming from
the dissociation of a strong acid and can be neglected. Similarly, when a strong base like NaOH dissolves in water, it dissociates
~100% into ions. Therefore, the [OH-] is equal to the molar concentration of the base.
Note
When a weak acid or a weak base dissolves in water, it partially dissociates into ions. Therefore, the [H3O+] or the [OH-] in the
cases of weak acids and weak bases has to be determined experimentally for the calculations.
Example 6.5.1
Formula: [ OH −
] =
10
[H3 O
+
]
−14
0.10
= 10
−13
M
Example 6.5.2
A vinegar solution has [H3O+] = 2.0 x 10-3. a) What is the hydroxide ion concentration in the vinegar solution? b) is the
solution acidic, basic, or neutral?
Solution
a) Given [H3O+] = 2.0 x 10-3. Desired [OH-] = ?
−14
Formula: [ OH −
] =
10
[H3 O
+
]
−14
Example 6.5.3
Calculations: [ H 3O
+
] = 10
−14
/0.010 = 10
−12
M
6.5.2 https://chem.libretexts.org/@go/page/371806
Example 6.5.4
a) Calculate the [H3O+] in an ammonia solution that has [OH-] = 4.0 x 10-4 M? b) Is the
solution acidic, basic, or neutral?
Solution
a) Given [OH-] = 4.0 x 10-4. Desired [H3O+] = ?
Formula:[ H 3O
+
] = Kw / [ OH
−
] = 10
−14
/ [ OH
−
]
−14
Calculations: [0H −
] =
10
−4
= 2.5 × 10
−11
M
4.0×10
6.5: Dissociation of water is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
6.5.3 https://chem.libretexts.org/@go/page/371806
6.6: The pH
What is pH?
The acidity of an aqueous solution relates to H3O+ ion concentration, but there are two problems with it. The first problem is that
the molar concentration of hydronium ion [H3O+] varies over a wide range, usually from 1-to-0.00000000000001 which is not easy
to comprehend. Log scale solves the problem, e.g., if a number increases from 1 to 1000, it increases from 0 to 3 on a log scale to
the base 10. The second problem is that the [H3O+] is usually a small number between 0 and 1. The log of a number that is between
0 and 1 is a negative number that is not easy to grasp mentally. If a number is between 0 and 1, its reciprocal is more than 1, and its
log is a +ve number.
The pH is defined as a log of reciprocal of the molar concentration of hydronium ions [H3O+]. It can also be stated that the pH
is a negative log of the molar concentration of hydronium ion [H3O+]. The mathematical form of the pH is the following.
1
+
pH = log = − log[ H3 O ]
+
[ H3 O ]
Calculating pH
First, determine the molar concentration of hydronium ion [H3O+]. Than take log base 10 of [H3O+] and change the sign of the
answer, i.e.:
+
pH = − log[ H3 O ]
Example 6.6.1
Calculate the pH of neutral water that has [H3O+] = 10-7M?
Solution
Given [H3O+] = 10-7 M, desired pH = ?
Formula: pH = − log[ H 3O
+
]
The keys-sequence for the calculation on a Microsoft Window’s scientific calculator is (the last box is the output of the keys-
sequence):
Example 6.6.2
Calculate the pH of the 0.010 M HCl solution?
Solution
Given [HCl] = 0.010 M = [H3O+], desired pH = ?
6.6.1 https://chem.libretexts.org/@go/page/371807
Formula: pH = − log[ H 3O
+
]
Note that the given number 0.010 M has two significant figures shown in bold fonts. So, the answer pH 2.00 also has two
significant figures shown in bold fonts.
Example 6.6.3
As the acid 0.01M and 0.10M is gradually added to increase the acid from neutral water in example 1 to 0.01M in example 2 to
0.10 M in example 3, , the pH gradually decreased from 7 to 2 to 1.
The pH of the acidic solution is less than 7. The more acidic the solution, the lower the pH.
The pH decreases as the [H3O+] increases because the reciprocal of [H3O+] is used in the pH calculation. The larger the given
number, the smaller the reciprocal; it translates to “the more acidic the solution, the lower the pH.
Example 6.6.4
Note that the given number 0.010 M has two significant figures shown in bold fonts. So, the answer pH 12.00 also has two
significant figures shown in bold fonts.
Example 6.6.5
Calculate the pH of 0.10 M NaOH solution?
Solution
Given [NaOH] = 0.10 M = [OH-], desired pH = ?
First, calculate [H3O+] from the give [OH-]: 10 −14
/ [ OH
−
] = 10
−14
/0.10 = 1.0 × 10
−13
6.6.2 https://chem.libretexts.org/@go/page/371807
As the base, 0.010M and 0.10M are gradually added in the previous two exams to the neutral water at pH 7, the pH increases from
7 to 12 to 13.
The pH of the basic solution is more than 7. The more basic the solution, the higher the pH.
pH measurement
The pH is usually measured in laboratories by digital pH meters. The electrode of the pH meter is first calibrated with solutions of
know pH values, and then the electrode is dipped in the test solution to read its pH value. Universal pH indicator papers are
available that turn to a specific color when placed in the solution. The pH is read by matching the color of the test paper with the
color on the chart. Fig. 6.6.1 shows a digital pH meter and two universal pH indicating papers commonly used for pH
measurements in laboratories.
Figure 6.6.1 : pH/Ion Meter (left) and universal pH indicator papers (middle and right). Source: Datamax / Public domain;
Bordercolliez / CC0; Fredquintao~commonswiki, CC-BY-SA-2.5.
The pH indicators are weak acids or weak bases that change color in a specific pH range. A few drops of the solution of a pH
indicator paper are added to the test solution in which pH is changed by adding acid or base. The change in the color of the solution
indicates the pH range in which the indicator changes the color. Figure 6.6.2 shows the colors and the color transition ranges of
some of the common pH indicators.
Figure 6.6.2 : The colors and the pH ranges of color transition of some of the common pH indicators. Source: OpenStax / CC BY
(https://creativecommons.org/licenses/by/4.0)
6.6.3 https://chem.libretexts.org/@go/page/371807
Calculating hydronium ion concentration from pH
The formula for calculating the molar concentration of hydronium ions [H3O+] is obtained by rearranging the pH formula.
pH = − log[ H O ], re-arranges to: [ H O ] = antilog(−pH ) = 10 .
+ + −pH
3 3
Example 6.6.6
The keys-sequence for the calculation on a Microsoft Window’s scientific calculator is (the last box is the output of the keys-
sequence):
The calculated answer in the above example is a lengthy number, but it is rounded to two significant figures shown in bold font
in 2.8×10 M , because the given number pH 7.56 has two significant figures shown in bold font.
−8
Significance of pH Scale
The pH scale varies from ~0 to ~14. The pH of 7 is neutral, pH more than 7 is basic, and pH less than 7 is acidic, as illustrated in
Fig. 6.6.3. Different foods have different pH values, as shown in Fig. 6.1.1. Similarly, several of the acids and bases in household
use have a specific pH range, as shown in Fig. 6.6.3.
Figure 6.6.3 : pH scale and pH of some common substances. pH 7 is neutral, pH less than 7 is acidic, and pH more than 7 is basic.
The control of pH is essential in the proper functioning of biological systems. Plants thrive if the soil does not have too acidic or
too basic a pH. Lacks and revers have a specific range of pH in which the aquatic life can survive. US natural water has a pH in the
6.6.4 https://chem.libretexts.org/@go/page/371807
range of 6.5 to 8.5 range. Seawater has a specific narrow range of pH 7.5 to pH 8.4 in which the life in the sea can function
correctly.
Figure 6.6.4 : The pH in various human body fluids is shown. Source: Modified from National Cancer Institute / Public domain
The blood has a pH of 7.4, and it can vary in a small range of 7.35 to 7.45. If the blood pH goes outside of the 7.35 to 7.45 range it
can result in medical problems. Enzymes in the body need a specific pH range because hydrogen bonding plays a vital role in the
structures needed for proper functioning. pH changes affect the hydrogen bonding and can make the enzymes less active or may
inactivate them. The pH of blood is maintained by a complex action of buffers that are described in the later sections. The pH of
urine can vary over a broad range from 4.6 to 8, depending on the recent diet and exercises.
Acid rain
Acidic gases like NO, NO2, N2O4, SO2 are released into the environment primarily during the combustion of fossil fuels. These
gases dissolve in the rainwater and make the rain acidic. For example, sulfur dioxide dissolves in water and makes sulfurous acid:
The acid rain damages the environment by making the soil, river water, and lack-water acidic. The acidic soil and water, in turn,
affect the plants and aquatic life. The acid water also reacts with calcium carbonate and corrodes metals that are responsible for
damage to the sculpture and other structures, as illustrated in Fig. 6.6.5, Fig. 6.6.6, and Fig. 6.6.7.
6.6.5 https://chem.libretexts.org/@go/page/371807
Figure 6.6.5 : Illustration of acid rain. Source: https://www.hiclipart.com/free-
trans...bacyp/download
Figure 6.6.7 : PEffect of acid rain on monuments. Source: Nino Barbieri / CC BY-SA
(http://creativecommons.org/licenses/by-sa/3.0/)
6.6: The pH is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
6.6.6 https://chem.libretexts.org/@go/page/371807
6.7: Acid-base reactions
Acid-base reactions are chemical reactions that involve the transfer of a proton (H+). Examples include the reactions of acids
with metals, carbonates, and Arrhenius bases, described in the following.
Figure 6.7.1 : Mg reacting with HCl and H2 gas bubbling out. Source: https://youtu.be/cpy_Zh-8sKA, CC BY
Alkali metals like Na give out one electron, alkaline earth metals like Mg give out two electrons, and aluminum gives out three
electrons. A proton can accept one electron. Therefore, the number of acidic protons should be equal to the number of electrons
given out to balance the electron transfer in these reactions.
Note
Noble metals or jewelry metals like gold, silver, platinum, copper are exceptions –they do not react with acids.
Example 6.7.1
write a balanced equation for the reaction of aluminum with HBr?
Solution
Step 1) write the equation with the correct formulas of reactants and products.
Note that Al gives out three electrons to make Al3+; that is why three Br- are attached to it in the product to balance the charge.
Step 2) balance the electrons lost by the metal with the number of acidic protons.
Step 3) Balance the rest of the elements by the hit and trial method. There are 3 hydrogen atoms on the left, balance them by
adding 3/2 coefficient to the hydrogen on the right.
3
Al(s) + 3HBr(aq) → AlBr3 (aq) + H2 ( g)
2
Step 4. Although the equation is balanced, it is recommended to remove the fraction by multiplying the coefficients in the
whole equation with the highest common factor to obtain the final balanced equation.
6.7.1 https://chem.libretexts.org/@go/page/371808
Reactions of acids with carbonates and hydrogen carbonates
Carbonic acid (H2CO3) is a weak acid found in carbonated water. The H2CO3 is a product of carbon dioxide (CO2) and water
(H2O) by the following equilibrium reaction.
CO (aq) + H O(l) −
↽⇀
− H CO (aq)
2 2 2 3
Hydrogen carbonate (HCO3-) and carbonate (CO32-) are one and two hydrogen less, respectively than carbonic acid. The salts
containing HCO3- and CO32- accept one and two protons, respectively, from acids to make H2CO3. The H2CO3 decomposes to
carbon dioxide and water by the reverse reaction shown above. Fig. 6.7.2 shows the reaction of sodium hydrogen carbonate with
hydrochloric acid.
Figure The reaction of sodium carbonate with HCl. Before and after HCl addition. Carbon dioxide
:
6.7.2
bubbles out after HCl addition. Source: NCSSM, 05/18/20, https://youtu.be/TJYOxGHNTzg, CC BY 3.0
Similarly, calcium carbonate found in limestone reacts with acids. For example, sulfuric acid is one of the components in acid rain
that reacts with calcium carbonate and damages sculptures made of stone.
CaCO3 ( s) + H2 SO 4 (aq) → CaSO4 (aq) + H2 O(l) + CO2 ( g) ↑
This is the net ionic equation of a reaction of an acid with an Arrhenius base. The cation of the base and the anion of the acid do not
react –they are spectator ions. The reaction equation for an acid-base reaction is written in various ways explained below.
Molecular equation
The molecular equation shows the formulas of the substances. For example, the molecular equation of the reaction of HCl with
NaOH is the following.
It is evident that Na+ and Cl- do not react –they are the spectator ions.
6.7.2 https://chem.libretexts.org/@go/page/371808
Net ionic equation
The spectator ions are present on both sides of the equation in equal numbers. The spectator ions can be canceled out, like the terms
in an algebraic equation.
+ − + − + −
H (aq) + Cl (aq) + Na (aq) + OH (aq) → H2 O(l) + Na (aq) + Cl (aq)
The net ionic equation shows the substances that are not spectator ions after canceling out the spectator ions from the complete
ionic equation. The net ionic equation of the reaction of HCl with NaOH is the following.
+ −
H (aq) + OH (aq) → H2 O(l)
Note
All of the reactions of the strong acids with strong Arrhenius bases have the same net ionic equation.
Writing a balanced chemical equation of the reaction of acids with Arrhenius bases
The following example shows the steps needed to write the balanced equation of the reaction of an acid with an Arrhenius base.
Example 6.7.2
Step 2) Balance the H+ in the acid with the OH- in the base.
Step 3) Balance the H2O with the H in the acid, or with the OH- in the base.
+
Step 4) Write the salt by combing the cations from the base and the anions from the acid. Make sure the charges are balanced
in the salt to make it a neutral substance.
2HCl(aq) + Ca(OH)2 (aq) → CaCl2 (aq) + 2 H2 O(I) (6.7.1)
Example 6.7.3
Step 2) Balance the H+ in the acid with the OH- in the base. They are already balanced in the above equation.
Step 3) Balance the H2O with the H+ in the acid, or with the OH- in the base.
Step 4) Write the salt by combing the cations from the base and the anions from the acid. Make sure the charges are balanced
in the salt to make it a neutral substance.
6.7.3 https://chem.libretexts.org/@go/page/371808
Antacids
The stomach sometimes produces excess HCl that may cause heartburn.
Antacids are the substances used to neutralize the excess HCl in the stomach.
Acid-base titration
The acid-base titration is an analytical process of determining the concentration of acid, called analyte, by neutralizing it with
a base of know concentration, called the standard, or vice versa.
Usually, the base is in a burette and added drop by drop to a known volume of the acid in an Erlenmeyer flask, as illustrated in Fig.
6.7.4. A few drops of an acid-base indicator are mixed with the acid. The indicator changes color at a specific narrow pH range.
The point when the stoichiometric amount of the base has been added to the acid is called the equivalence point.
The point when the indicator changes color is the end-point.
6.7.4 https://chem.libretexts.org/@go/page/371808
Figure 6.7.4 : Acid-base titration set-up. The analyte color due to the indicator before, at the endpoint, and after being shown by the
images on the left, middle, and right, respectively. Source: Taken from Luigi Chiesa / Public domain.
The calculation steps after the titration are usually the following.
Example 6.7.4
What is the molarity of the HCl solution if 50.0 mL of the HCl solution requires 42.0 mL of 0.123 M NaOH solution in the
titration?
Solution
Step 1) Given:
0.123 mol N aOH
[NaOH] = 0.123 M =
1 NaOH
1 L NaOH
Volume of standard = 42.0 mL. = 42.0 mL NaOH × = 0.0420 L NaOH
1000 mL NaOH
1 L HCl
Vol. analyte = 50.0 mL = 50.0 mL HCl × = 0.0500 L HCl
1000 mL HCl
Step 2) Calculate the moles of the standard from the volume and the molarity product.
0.123 mol NaOH
Moles of NaOH~consumed = 0.0420 L NaOH × = 0.00517 mol NaOH
1 L NaOH
Step 4) write the conversion factor for mole of standard to mole of analyte calculation from the equation.
6.7.5 https://chem.libretexts.org/@go/page/371808
1mol HCl
1 mol NaOH
Step 5) Calculate the moles of the analyte by multiplying the moles of the standard and the conversion factor.
1 mol HCl
0.00517 mol NaOH × = 0.00517 mol HCl
1 mol NaOH
Step 6) Calculate the molarity of the analyte by dividing the moles with the volume in liters of the analyte.
1 mol HCl
0.00517 mol NaOH × = 0.00517 mol HCl
1 mol NaOH
After learning the steps thoroughly, the calculations can be done in a few steps, as shown in the following.
Calculations:
1 L NaOH 0.123 mol NaOH 1 mol HCl
42.0 mL NaOH × × × = 0.00517 mol HCl
1000 mL NaOH 1 L NaOH 1 mol NaOH
Molarity of HCl:
n(mol) 0.00517 mol HCl 1000 mL HCl
M = = × = 0.103 M HCl
V (L) 50.0 mL HCl 1 L HCl
Example 6.7.5
What is the molarity of the H2SO4 solution if 50.0 mL of the H2SO4 solution requires 32.3 mL of 0.201 M NaOH solution in
the titration?
Solution
Step 1) Given:
0.201 mol NaOH
[NaOH] = 0.201 M =
1 L NaOH
1 L NaOH
Vol. standard = 42.0 mL = 32.3 mL NaOH × = 0.0323 L NaOH
1000 mL NaOH
1 L HCl
Vol. analyte = 50.0 mL = 50.0 mL HCl × = 0.0500 L HCl
1000 mL HCl
Step 2) Calculate the moles of the standard from the volume and the molarity product.
0.201 mol NaOH
Moles of NaOH consumed = 0.0323 N NaOH × = 0.00649~mol N aOH
1 L NaOH
Step 4) write the conversion factor for the moles of standard to mole of analyte calculation from the equation.
1 mol H 2 SO 4
2 mol NaOH
Step 5) Calculate the moles of the analyte by multiplying the moles of the standard and the conversion factor.
1 mol H 2 SO 4
0.00649 mol NaOH × = 0.00325 mol H 2 SO4
2 mol NaOH
Step 6) Calculate the molarity of the analyte by dividing the moles with the volume in liters of the analyte.
0.00325 mol H 2 SO 4
Concentration of H 2 SO 4 = = 0.0649 M H2 S O4
0.0500 L H 2 SO 4
6.7.6 https://chem.libretexts.org/@go/page/371808
The same calculation in a summary form is in the following.
0.00325 mol H 2 SO 4
Concentration of H 2 SO 4 = = 0.0649 M H 2 SO 4
0.0500 L H 2 SO 4
= 0.00326 mol H 2 SO 4
6.7: Acid-base reactions is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
6.7.7 https://chem.libretexts.org/@go/page/371808
6.8: pH Buffers
What is a pH buffer?
A pH buffer is an aqueous solution consisting of a weak acid and its conjugate base or vice versa that lets the pH change to be
minimal when a small amount of a strong acid or a strong base adds to it.
For example, the addition of 0.020 mol HCl into 1 L of water changes pH from 7 to 1.7, i.e., about an 80% change in pH. The
addition of 0.020 mol NaOH to the same water changes pH from 7 to 12.3, i.e., again, about an 80% change in pH. In contrast to
pure water, 1 L of buffer solution containing 0.50 mol acetic acid (CH3COOH) and 0.50 mol CH3COO- -the conjugate of the acetic
acid, changes pH from 4.74 to 4.70 by the addition of the same 0.020 mol HCl and from 4.74 to 4.77 by the addition of 0.020 mol
NaOH, i.e., about 1% change in pH, as illustrated in Fig. 6.8.1.
Figure 6.8.1 : Effect acid and base addition on pH change of pure water at pH 7.00 and on acetic acid/sodium acetate buffer at pH
4.74.
Buffers of different initial pH values can be prepared, e.g., by varying the ratio of the weak acid to its conjugate base or by using a
different set of a weak acid and its conjugate base. One example is a buffer of initial pH 4.74 comprising 0.5 M acetic acid and 0.5
M sodium acetate shown in Fig. 6.8.1. Another example is a buffer comprising 0.1M dihydrogen phosphate and 0.1M hydrogen
phosphate that has an initial pH of 7.21, as shown in Fig. 6.8.2.
6.8.1 https://chem.libretexts.org/@go/page/372205
Figure 6.8.2 : Effect of acid and base addition on pH change of pure water at pH 7.00 and on Na2HPO4/NaH2PO4 buffer of pH
7.21.
Buffer capacity
Buffer capacity refers to how much amount of a strong acid or a strong base the buffer can handle before a drastic change in
pH happens.
Higher the amount of weak acid/conjugate base higher the buffer capacity. The buffer has an equal amount of weak acid and its
conjugate base has a higher buffer capacity than the same buffer that has an unequal ratio of the acid and its conjugate base.
Caution
Strong acid and its conjugate base or a strong base and its conjugate acid do not make a buffer solution.
The molar concentration of hydrogen ions [H+] defines the pH of the solution. The conjugate base consumes any strong acid added.
− −
HA + CH3 COO → CH3 COOH + A
, where HA is any strong acid and A- is its conjugate base. The concentration of CH3COOH increases and CH3COO- decrease, but
pH decreases small because [H+] increases only a little. Similarly, the weak acid consumes any strong base added.
− +
MOH + CH3 COOH → CH3 COO +M + H2 O
Where MOH is any strong base, and M+ is its conjugate acid. Fig. 6.8.3 illustrates the mechanism of buffer action.
6.8.2 https://chem.libretexts.org/@go/page/372205
Figure 6.8.3 : Mechanism of blood buffer action: The main blood buffer is H2CO3/HCO3-. Added acid is consumed by reacting
with the HCO3- and added base is consumed by consumed by reacting with H2CO3, maintaining the pH ~7.4. Longs regulate the
pH by regulating the rate of CO2 exhale and kidney regulates the pH by regulating the HCO3- in the buffer system.
Note
Strong acid and its conjugate base mixture have no buffer action because the acid dissociates almost 100%, leaving no acid
behind for the buffer action. The conjugate base of a strong acid is a very weak base that does not react with the added acids.
The same explanation applies to a strong base and its conjugate acid mixture having no buffer action.
Figure 6.8.4 : Mechanism of blood buffer action: The main blood buffer is H2CO3/HCO3-. Added acid is consumed by reacting
with the HCO3- and the added base is consumed by reacting with H2CO3, maintaining the pH ~7.4. Longs regulate the pH by
regulating the rate of CO2 exhale and the kidney regulates the pH by regulating the HCO3- in the buffer system.
6.8.3 https://chem.libretexts.org/@go/page/372205
The primary mechanism for pH regulation by the H2CO3/HCO3- buffer in the blood is through the lungs. When the blood pH is
acidic compared to the average, the breathing rate increases exhaling more CO2 that decreases the concentration of H+, following
the blue arrows in Fig. 6.8.4, increasing the pH. The decrease in the breathing rate has the opposite effect. Kidneys also regulate the
blood pH by adding or removing HCO3-, but the kidney’s response is delayed compared to the response of the lungs.
6.8: pH Buffers is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
6.8.4 https://chem.libretexts.org/@go/page/372205
CHAPTER OVERVIEW
7: Gases
7.1: Characteristics of gases
7.2: The pressure-volume relationship
7.3: The temperature-volume relationship
7.4: The pressure-temperature relationship
7.5: The combined gas law
7.6: The volume-amount relationship
7.7: Ideal gas law
7.8: Dalton’s law of partial pressure
7: Gases is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
1
7.1: Characteristics of gases
What is a gas
Gas is one of the four states of matter that falls between the liquid and the plasma state. The air around us is a gas composed of
~78% nitrogen (N2), ~21% oxygen, and the remaining ~1% are other gases, including carbon dioxide, water vapors, argon, etc., as
illustrated in Fig. 7.1.1.
Figure 7.1.1 : Depictions of the composition of the Earth's atmosphere; the lower pie represents the least common gasses that
compose 0.037680% of the atmosphere. Source: Brockert at English Wikipedia / Public domain
Elements that exist as gases at room temperature and atmospheric pressure include noble gases which are monoatomic molecules
including helium (He), neon (Ne), argon (Ar), krypton (Kr), xenon (Xe), and radon (Rn), and diatomic elemental molecules that
include hydrogen (H2), nitrogen (N2), oxygen (O2), and two halogens: fluorine (F2) and chlorine (Cl2). Several molecular
compounds exist as gases at room temperature, e.g., carbon dioxide (CO2), carbon monoxide (CO), methane (CH4), nitrogen
dioxide (NO2), sulfur dioxide (SO2), hydrogen sulfide (H2S), ammonia (NH3).
The properties of gases are defined by their amount in moles (n), volume (V), temperature (T), and pressure (P). These parameters
are described in the following.
7.1.1 https://chem.libretexts.org/@go/page/371599
Note
The amount of gas in moles is needed in gas laws because the properties of gases are proportional to the number of molecules,
not to the masses of the molecules.
Example 7.1.1
Step 4. Multiply the given quantity with the conversion factor that cancels the given and leaves the desired unit in the answer:
1 mol O2
10.0 g O2 ) × = 0.312 g O2
32.00 g O2
Usually, the volume is reported in liters (L), which is dm3, or milliliter (mL), which is cm3. The relationship between the volume
units is the following.
3
1 m = 1000 L
1 L = 1000 mL
The gases do not have a fixed shape or volume. The gases acquire the shape of the container. The gases expand or contract to
fill the available space in a container.
The temperature is a manifestation of the thermal energy of a substance. Thermal energy is a source of heat. Heat is the flow of
energy from a hot to a cold object. The SI unit of temperature is Kelvin (K) which has a value of 273.15 K at the freezing point of
water and 375.15 K at the boiling point of water. The zero of the kelvin scale is called the absolute zero at which no more energy
can flow out of a substance as heat. There is no negative number on the kelvin scale of temperature.
Other commonly used units of temperature are Celsius (oC) which has a value of 0 oC at the freezing point of water and 100 oC at
the boiling point of water. The relationship between temperature in kelvin (TK) and the temperature in Celsius (TC) is the following.
TK = TC + 273.15
Fahrenheit is a temperature scale used in the English system of measurement that has a value of 32 oF at the freezing point of water
and 212 oF at the boiling point of water. The relationship between the temperature in Fahrenheit (TF) and the temperature in Celsius
(TC) is the following.
9
TF = TC + 32
5
7.1.2 https://chem.libretexts.org/@go/page/371599
Note
A
.
The matter has mass and applies a force, which is the weight due to gravitational pull. For example, a column 1 m x 1m at sea level
and height equal to the earth’s atmosphere, as illustrated in Fig. 7.1.2 has a mass of ~10,000 kg. Mercury (Hg) filled in a column of
1 m x 1m and height 0.760 m has the same mass as the mass of air of 1 m x 1 m column extending from the sea level to the entire
atmosphere of the earth.
Figure 7.1.2 : Illustration of an air column of 1m x 1m base and height covering the entire atmosphere of the earth. Modified from:
http://wikimapia.org/8706099/Norfolk.../photo/1606686
A column of air extending from sea level to the entire atmosphere applied pressure equal to one atmosphere (atm), where atm is the
pressure unit. One atmosphere (atm) is equal to 760 millimeter of mercury (mmHg), where mmHg is another unit of pressure. One
mmHg is also called Torr. The relationship between atm, mmHg, and Torr is the following.
Pascal (Pa)
The SI unit of pressure is pascal (Pa), which is pressure (P) applied by a force (F) of one newton (N) on an area of one meter
square (m2). Mathematical form of the unit is: 1 P ascal = or P a =
1 N ewton
1 meter−square
N
m2
Kilopascal is 1000 Pa, and kilopascal is also called the bar. The English system unit of pressure is pound-force per square inch
(psi). The relationship between different units of pressure mentioned above is the following.
5 2
1 atm = 760 mmHg = 760 Torr = 1.01325 × 10 Pa = 1.01325 × 10 kP = 1.01325 bar = 14.7 psi
These relationships are changed to conversion factors that allow conversion from one to another pressure unit, as explained in the
following examples.
Example 7.1.2
760 mmHg
and 1 atm
Calculations: multiply the given quantity with the conversion factor that cancels the given unit and leaves the desired unit in
the answer.
7.1.3 https://chem.libretexts.org/@go/page/371599
1 atm −2
51 mmHg × = 6.7 × 10 atm
760 mmHg
5
760 mmHg
The conversion factors from the equality: 5
and 1.01325×10 Pa
760 mmHg
1.01325×10 Pa
Calculations: multiply the given quantity with the conversion factor that cancels the given unit and leaves the desired unit in
the answer.
5
1.01325 × 10 Pa 3
51 mmHg × = 6.8 × 10 Pa
760 mmHg
The barometer, shown in Fig. 7.1.3, was invented by Evangelista Torricelli. It is a glass tube that is more than 760 mm long and
closed at one end. The glass tube is filled with mercury and inverted to dip its open end in a dish that contains a layer of mercury.
The mercury falls from the tube due to its weight but stops falling when the column of mercury is 760 mm in height at sea level due
to the atmospheric pressure pushing it back into the tube. Remember 760 mmHg = 1 atm. Above 760 mm is a vacuum in the tube.
The pressure of gases in a laboratory is usually measured using an instrument called a manometer.
A manometer is a U-shaped glass tube partially filled with mercury, as shown in Fig. 7.1.4. One end of the U-tube is connected to a
gas chamber in which pressure is being measured, and the other end is either closed (with a vacuum on the closed end), called a
closed-ended manometer, or it is open to the atmosphere, called an open-ended manometer.
7.1.4 https://chem.libretexts.org/@go/page/371599
Figure 7.1.4 : Manometers. Source: Modified from Algarabia / Public domain
In the case of a closed-ended manometer, the gas pressure is equal to the difference in the height of mercury in the two arms of
the U-tube in mmHg, i.e., height (h) from point N to point B in mm units as shown in Fig. 7.1.4.
Example 7.1.3
Calculate the pressure in atm for the gas shown in Fig. 7.1.4 in the closed-ended manometer if h = 40 mm?
Solution
In a closed-ended manometer, the gas pressure is equal to the height h in mm Hg: P = 40 mm Hg.
Pressure in atm:
1 atm
−2
40 mmHg × = 5.3 × 10 atm
760 mmHg
In the case of an open-ended manometer, the gas pressure is equal to the difference in the height of mercury in the two arms of
the U-tube in mm of Hg plus the atmospheric pressure outside expressed in mm of Hg.
Figure 7.1.5 : Air pressure decreases and the air concentration and oxygen concentration in it decreases as the altitude increases.
Source: Brendan Scott/The Conversation, CC BY-ND https://images.theconversation.com/f....png?ixlib=rb-
Blood pressure
The heart pumps the blood into the circulatory system. When the heart contracts, it applies the pressure on the blood in it, and the
blood pumps out of the heart into the circulatory system, as illustrated in Fig. 7.1.6. The blood pressure in the circulatory system is
highest at this point, and it is called systolic pressure. It can be in the range of 100 to 200 mm Hg. A desirable systolic pressure
range is 100 to 120 mm Hg. When heart muscles relax, the cavity in the heart expands and more blood fills in the heart. The blood
7.1.5 https://chem.libretexts.org/@go/page/371599
pressure in the circulatory system is minimum at this point, and it is called diastolic pressure. The diastolic pressure may vary in
the range of 60 to 110 mm Hg. A desirable diastolic pressure is less than 80 mm Hg.
Figure 7.1.6 : Diagram of heart illustrating systole, i.e., blood pumping out of heart vs Diastole i.e., blood filling into the heart.
Source: BruceBlaus / CC BY-SA (https://creativecommons.org/licenses/by-sa/4.0)
The device used to measure blood pressure is called a sphygmometer. It consists of a cuff that wraps around the upper arm, a
pump to inflate air in the cuff, and a stethoscope to hear the sound of blood flow, as illustrated in Fig. 7.1.7. The cuff is inflated
with air to a pressure above the systolic pressure and it results in cutting off the blood flow through the brachial artery in the upper
arm. A stethoscope is used to hear the sound of blood flow. No sound is heard at this point. The pressure in the cuff is slowly
reduced. When the pressure in the cuff is equal to systolic pressure, the blood begins to spurt into the artery, and a tapping sound is
heard through the stethoscope. At the point when the pressure is equal to the diastolic pressure, the blood flows freely through the
artery, and the tapping sound disappears. The blood pressure reading is reported as a set of two numbers, e.g., 120/80, where the
higher number is systolic, and the lower number is the diastolic pressure.
Figure 7.1.7 : A healthcare professional performing blood pressure monitoring on a patient. Source: Roa'a najim / CC BY-
SA (https://creativecommons.org/licenses/by-sa/3.0)
7.1.6 https://chem.libretexts.org/@go/page/371599
Figure 7.1.8 : A model of a gas that is based on kinetic molecular theory of gases. The molecules are the spheres and the tails
attached to the molecules represent the trajectories of molecular motion. The red tails represent molecules bouncing back from the
wall resulting in the pressure exerted by the gas on the wall of the container. Source: Becarlson / CC BY-SA
(https://creativecommons.org/licenses/by-sa/3.0)
Explanation of the characteristics of the gases based on the kinetic molecular theory of gases
1. Upon exposing gas to a vacuum, the gas molecules keep moving in a straight path in the vacuum until they collide with a
container boundary and bounce back. It explains gases are expandable. Collisions of gas molecules with the wall exert outward
pressure on the wall and, as a reaction, the wall exerts an equal inward pressure on the gas. Applying more pressure to the gas
through any portion of the gas boundary disturbs the above-mentioned balance of action and reaction forces, resulting in an
inward movement of the portion of the boundary surface. In other words, the gases are compressible.
2. If a boundary separating two gases is removed, the random molecular motions ultimately mix the two gases making a
homogeneous mixture, i.e., gases are entirely miscible with other gases.
3. When temperature increases the gas molecules move faster and colloid more frequently, which explains the fact that gases exert
more pressure upon heating and expand if the boundary is not rigid. The more frequent bursting of tires in summer is due to the
pressure increase due to heating.
4. When more gas fills in the tire, the pressure increases due to more frequent collisions. Compressing the gas into smaller volume
have the same effect, i.e., the molecular collisions increase increasing the pressure.
The physical characteristics of the gases are independent of the nature of the gas molecules due to negligible intermolecular
interactions but depend on four physical properties, i.e., pressure, volume, temperature, and amount of gas. The following sections
describe the relationships between the physical properties of the gases.
7.1: Characteristics of gases is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
7.1.7 https://chem.libretexts.org/@go/page/371599
7.2: The pressure-volume relationship
Pressure-volume relationship
Consider a gas in a cylinder with a piston pushing it down due to massive objects placed on it and due to outside air pressure, as
illustrated in Fig. 7.2.1, The gas molecules strike the piston surface, applying pressure upward equal to the downward pressure
applied by the piston that keeps the piston stationary. Increasing the pressure on the piston, e.g., by adding more weight to it, causes
the piston to move down, reducing the gas volume. The gas molecules have less distance to travel before striking the piston surface
which increases the collision frequency and causes the gas pressure to increase until becomes equal to the outside pressure.
Figure 7.2.1 : Illustration of Boyle’s law: for a given amount of gas at a constant temperature, the volume is inversely proportional
to the pressure. Source: NASA's Glenn Research Center / Public domain.
When a change in one parameter causes a change in another, the parameters are related. When an increase in one parameter causes
a decrease in another, the two are inversely proportional to each other. Robert Boyle studied the quantitative relationship between
the volume and pressure of the gas, keeping the quantity of gas and the temperature constant. The research concluded in a law
called Boyle’s law, which states that:
Boyle's law
The volume of a gas is inversely proportional to the pressure of the gas provided the temperature and the amount of the gas
remain constant.
Figure 7.2.2 : Boyles law: the volume of the gas is inversely proportional to the pressure of the gas. Source: Robert Boyle, A
Defence of the Doctrine Touching the Spring And Weight of the Air (1662)
Fig. 7.2.2 illustrates the results of Boyle’s experiment. The mathematical form of Boyle’s law is:
1
V ∝
P
, or
k
V =
P
, or
PV = k
, where k is a consant. The graph of volume versus pressure is curvilinear, but the graph of volume versus the reciprocal of pressure
is a linear graph showing the inverse proportionality between the volume and the pressure. Since the product PV is a constant, it
7.2.1 https://chem.libretexts.org/@go/page/371600
implies that:
P1 V1 = P2 V2 = k
, i.e., a product of initial pressure (P1) and initial volume (V1) is equal to the product of final pressure (P2) and final volume (V2) of
gas provided the quantity of the gas and temperature does not change.
Example 7.2.1
The pressure of a 1.32 L sample of SO2 gas at 0.532 atm is increased to 1.231 atm. Calculate the new volume of the gas if the
temperature and the quantity of the gas remain the same?
Solution
Given: P1 = 0.532 atm, P2 = 1.231 atm, V1 = 1.32 L V2 = ?
P1 V1
Formula: P 1 V1 = P2 V2 = k , rearrange to isolate the desired variable: V 2 =
P2
.
1.231 atm
= 0.570 L
Example 7.2.2
An oxygen tank holds 20.0 L of oxygen at a pressure of 10.0 atm. What is the final pressure when the gas is released and
occupies a volume of 200 L?
Solution
Given: V1 = 20.0 L, V2 = 200 L, P1 = 10.0 atm P2 = ?
P1 V1
Formula: P 1 V1 = P2 V2 = k , rearrange to isolate the desired variable: P 2 =
V2
.
200 L
= 1.00 atm .
Breathing process
Boyle’s law explains the mechanism of the breathing process. Lungs are elastic structures like balloons placed in the thoracic
cavity, as illustrated in Fig. 7.2.3. The diaphragm muscle makes a flexible floor and ribs surround the cavity.
Figure 7.2.3 : Illustration of breathing process –inspiration and expiration occur due to the expansion and contraction of the thoracic
cavity, respectively. Source: OpenStax College / CC BY (https://creativecommons.org/licenses/by/3.0)
Inhalation
The inhalation or the inspiration process starts when the diaphragm contract and move down and the rib muscles contract,
expanding the thoracic cavity. Volume increases, the air pressure decreases inside the inside thoracic cavity and the atmospheric air
flows into the lungs until the pressure in the lungs is equal to the outside pressure.
7.2.2 https://chem.libretexts.org/@go/page/371600
Exhalation
The exhalation or the expiration process starts when the diaphragm expands and moves upwards, and the rib muscles relax,
contracting the thoracic cavity. Volume decreases and the air pressure increases inside the thoracic cavity that pumps the air out of
the lungs into the atmosphere.
7.2: The pressure-volume relationship is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif
Malik.
7.2.3 https://chem.libretexts.org/@go/page/371600
7.3: The temperature-volume relationship
Consider a gas in a cylinder with a piston in Fig. 7.3.1. Increasing temperature increases the average kinetic energy (KE) of the gas
molecules. The kinetic energy (KE) is directly proportional to the velocity of the molecules, i.e., KE = mv , where m is the 1
2
2
mass and v is the velocity. So, increasing temperature increases the velocity resulting in more frequent and more forceful collisions
resulting in increased gas pressure inside the chamber. The gas volume starts to increase causing the pressure to decrease until the
pressure inside the chamber is equal to the pressure outside. In other words, increasing temperature increases the volume of the gas
if the pressure and amount of gas are not changed.
Figure 7.3.1 : Increasing temperature increases the volume of the gas i.e., . Source: NASA's Glenn Research
V1 V2
= = k
T2 T2
Charles's law
Charles’s law states that the volume of a given amount of gas is directly proportional to the temperature in the Kelvin scale at
constant pressure.
Fig. 7.3.2 demonstrates that the volume of a gas decreases when the gas is cooled down.
Figure 7.3.1 : Air balloon (left) shrinks when its temperature is decreased by submerging in liquid nitrogen (middle) and re-expands
when returned to room temperature condition (right). Source: Ryan Poling aka expictura on
Flickr.https://en.Wikipedia.org/wiki/File:Nitrogen.ogv#file, CC BY 2.0
The mathematical forms of Charles’s law are the following.
V ∝T
, or
V = kT
, or
V
=k
T
, where k is a constant, V is volume, and T is the temperature (in kelvin scale) of the gas. Since V
T
is a constant, it implies that
V1 V2
= =k
T1 T2
where V1 is the initial volume, T1 is the initial temperature in Kelvin, V2 is the final volume, and T2 is the final temperature in
Kelvin, provided the amount of gas and pressure do not change. Note that the kelvin scale is used in Charles’s law because the
7.3.1 https://chem.libretexts.org/@go/page/371601
kelvin scale does not have negative numbers which means the linear curve starts from the origin without any y-intercept. If the
given temperature is not in the kelvin scale, first convert the temperature to the Kelvin scale and then use the gas laws for the
calculations.
Example 7.3.1
A sample of CO2 occupies 3.23 L volume at 25.0 oC. Calculate the volume of the gas at 50.0 oC if pressure and amount of gas
do not change?
Solution
Given: T1 = 25.0 oC + 273 = 298 K, T2 = 50.0oC + 273 = 320 K, V1 = 3.23 L, V2 = ?
V1 V2 V1 T1
Formula: T2
=
T2
, rearrange the formula to isolate the desired variable: V 2 =
T2
273 K
= 3.79 L
Charles’s law explains the drifting of warm air upward in the atmosphere. As the gas is wormed, its volume increases and its
density decreases which makes the gas drift upward. A hot air balloon, shown in Fig. 7.3.3 operates using hot air.
Figure 7.3.3 : A hot air balloon seen from a view directly below. The burner, or flame, is firing into the envelope above. The warm
air is less dense than the atmospheric air which makes the balloon rise in the air. Source: Arpingstone at English Wikipedia / Public
domain.
7.3: The temperature-volume relationship is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif
Malik.
7.3.2 https://chem.libretexts.org/@go/page/371601
7.4: The pressure-temperature relationship
Consider a gas in a cylinder with a piston in Figure 7.4.1. Increasing temperature increases the average kinetic energy (KE) and the
average velocity of the gas molecules resulting in more frequent and more forceful collisions which result in increased gas pressure
applied on the piston or the walls of the gas container.
Gay-Lussac's law
Gay-Lussac’s law states that the pressure of a gas is directly proportional to the absolute temperature provided the volume and
amount of gas are not changed.
Figure 7.4.1 : Increasing temperature increases pressure, i.e., . Source: NASA's Glenn Research Center / Public
P1 P2
=
T1 T2
domain.
The mathematical forms of Gay-Lussac’s law are the following.
P ∝T
, or
P = kT
, or
P
= k,
T
where k is a constant, P is pressure, and T is the temperature (in kelvin scale) of the gas. Since Since P
T
is a constant, it implies
that
P1 P2
= = k,
T1 T2
Example 7.4.1
The pressure of an oxygen tank containing 15.0 L oxygen is 965 Torr at 55 oC. What will be the pressure when the tank is
cooled to 16 oC.
Solution
First, convert the temperatures to the Kelvin scale before applying gas laws.
Given: T1 = 55 oC + 273.15 = 328.15 K, T2 = 16 oC + 273.15 = 289.15 K, P1 = 965 Torr, P2 = ?
Formula:
P1 P2
= ,
T1 T2
7.4.1 https://chem.libretexts.org/@go/page/371602
P1 T2
P2 =
T1
7.4: The pressure-temperature relationship is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad
Arif Malik.
7.4.2 https://chem.libretexts.org/@go/page/371602
7.5: The combined gas law
The laws relating to pressure P , volume V , and temperature T for a constant amount n of a gas are the following:
1. If n and T are constant: P 1 V1 = P2 V2 , that is Boyle's law.
V1 V2
2. If P and n are constant: = , that is Charles's law.
T1 T2
P1 P2
3. If V and n are constant: = , that is Gay Lussac's law.
T1 T2
The combined gas law allows calculating the effect of varying two parameters on the third.
Example 7.5.1
A weather balloon contains 212 L of helium at 25 C and 750 mmHg. What is the volume of the balloon when it ascends to an
∘
altitude where the temperature is −40 C and 540 mmHg, assuming the quantity of gas remains the same?
∘
Solution
Given and desired parameters (temperatures must be converted to Kelvin scale):
∘
P1 = 750 mmHg, V1 = 212 L, T 1 = 25 C + 273.15 = 298.15 K
∘
P2 = 540 mmHg, V2 =? T 2 = −40 C + 273.15 = 233.15 K
Formula:
P1 V1 P2 V2
= ,
T1 T2
Calculations:
7.5: The combined gas law is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
7.5.1 https://chem.libretexts.org/@go/page/371603
7.6: The volume-amount relationship
Figure 7.6.1 illustrates the effect of the amount of gas on the volume. Adding more gas molecules increases the collision frequency
of the molecules with the walls increasing the gas pressure. The gas expands to reduce pressure until the pressure of gas in the
chamber is equal to the outside pressure.
Figure 7.6.1 : When more gas is added the volume of a gas increases, i.e., V1/n1 = V2/n2. Source: NASA's Glenn Research
Center / Public domain.
Avogadro's law
Avogadro’s law states that the volume of a gas is directly proportional to the amount of gas in moles provided the temperature
and pressure of the gas are not changed.
V ∝n
or
V = kn
or
V
=k
n
where V is the volume, n is the number of moles, and k is the constant for the gas under the conditions of constant temperature and
pressure. Since V/n is a constant, it implies that:
V1 V2
= =k
n1 n2
where V and n is initial volume and the initial amount of gas, respectively, and V and n is final volume and the final amount of
1 1 2 2
gas in mole, provided the temperature and pressure are not changed.
Example 7.6.1
A weather balloon containing 3.0 moles of helium has a volume of . What is the final volume if
66 L 2.0 moles of helium is
added to it. The pressure and temperature of the gas do not change?
Solution
Given V1 = 66L, n1 = 3.0 mol, V2 =?, n2 = 3.0 + 2.0 = 5.0 mol
Formula:
V1 V2
= ,
n1 n2
7.6.1 https://chem.libretexts.org/@go/page/371604
V1 n2
V2 = .
n1
7.6: The volume-amount relationship is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif
Malik.
7.6.2 https://chem.libretexts.org/@go/page/371604
7.7: Ideal gas law
Combined relationships between pressure, volume, temperature, and amount of gases
The gas laws described above give the following relationships between volume V , pressure P , the temperature in kelvin T , and the
amount of gas in moles n :
V ∝ 1/P , at constant T and n ,
V ∝T , at constant P and n , and
V ∝n , at constat T and P .
The three proportionalities combine to give the following ideal gas relationship:
nT
V ∝
P
, that rearranges to
P V = nRT
pressure that can be in various units, and V is the volume that can be in various units. The value of R in different units of P , V ,
and P V products are given in Table 1. If values of any three among the P , V , n, and T are known, the value of the fourth one can
be calculated by using the ideal gas law.
Caution
In these calculations, the units of R should be in agreement with the units of P , V , n, and T . If they are not in agreement, the
given unit of P , V , n, and T must be converted to agree with the units of R .
0.08206 L-atm/mol-K
8.314 J/mol-K
1.987 cal/mol-K
8.314 m3-Pa/mol-K
62.36 L-torr/mol-K
nT
=R is a constant, that implies that:
P1 V1 P2 V2
= =R
n1 T1 n2 T2
7.7.1 https://chem.libretexts.org/@go/page/371605
P1 V1 P2 V2
If one or two parameters in the ideal gas equation n1 T1
=
n2 T2
are constant; they cancel out, leaving the relationship between the
remaining parameters, e.g.
if n and T are constant: P 1 V1 = P2 V2 , that is Boyle's law,
V1 V2
if P and n are constant: T1
=
T2
, that is Charles's law,
V1 V2
if P and T are constant: n1
=
n2
, that is Avogadro's law,
P1 P2
if V and n are constant: T1
=
T2
, that is Gay Lussac's law,
P1 P2
if V and T are constant: n1
=
n2
, that is pressure-mole relationship,
P1 V1 P2 V2
and if n is constant: T1
=
T2
, that is combined gas law.
Example 7.7.1
Calculate the volume of 1.000 mole of a gas in liters (L) at 0.000 oC and 1.000 atm?
Solution: Given n =1.000 mol, T = 0 oC+273.15=273.15 K, P =1.000 atm, and R =0.08206 L-atm/mol-K
Solution
Given n =1.000 mol, T = 0 oC+273.15=273.15 K, P =1.000 atm, and R =0.08206 L-atm/mol-K
Formula: P V = nRT , rearrange the formula to: V =
nRT
P
.
L−at m
1.000 mol×0.08206 ×273.15 K
Note
The volume of 1 mol of an ideal gas is 22.41 L at 0 oC and 1 atm pressure, as calculated in the above example
Example 7.7.2
Calculate the volume of a container that has 1.50 mol of He gas at 7.2 atm and 25 oC?
Solution
P =7.2 atm, V =? n =1.50 mol, T =25 oC}+273=298 K, R=0.08206 L − atm/mol − K
Formula: P V = nRT , rearrange the formula to: V =
nRT
L−at m
1.50 mol×0.08206 ×298 K
Example 7.7.3
Calculate the pressure in a 5.1 L container that has 0.60 mol of He at 25 oC?
Solution
= ?, V =5.1 L, n = 0.60 mol, T =25 oC+273=298 K, R = 0.08206
L−atm
P
mol−K
Formula: P V = nRT , rearrange the formula, plug in the values and calculate:
L−at m
0.60 mol×0.08206 ×298 K
nRT m ol−K
P = = = 2.9 atm
V 5.1 L
7.7.2 https://chem.libretexts.org/@go/page/371605
Example 7.7.4
Calculate the pressure of 0.60 mol of He in example 3 mixed with 1.50 mol of He in example 2 in a container of 5.1 L volume
at 25 oC?Solution
Ptotal = ?, n total = 1.5 mol + 0.60 mol = 2.1 mol, V = 5.1 L, T = 25 oC + 273 = 298 K, and R = 0.08206 L−atm
mol−K
Note
The P of 2.1 mol He in example 4 is equal to P of 1.50 mole He in example 2+P of 0.6 mol He in example 3, i.e., 7.2 atm
total
+2.9 atm =10.1 atm. This calculation demonstrates that when gases are mixed, the total pressure is the sum of the pressures that
each fraction will exert if it was alone in that space. It is demonstrated by mixing the same gas, i.e., He with He, but it remains
true when different gases are mixed, as long as all the gases involved obey the ideal gas law.
Currently accepted STP is 0oC and 1 bar. The molar volume of an ideal gas at 0oC and 1 bar is 22.71 L, but for most practical
purposes, the older definition of STP of 0oC and 1 atm is used.
The calculations in example 1 of the previous section show that the molar volume of an ideal gas is 22.41 L at STP. Fig. 7.7.1
illustrates the molar volume of an ideal gas at STP.
Figure 7.7.1 : The molar volume of gases at STP is 22.4 L which is the volume of 11.1” x 11.1” x 11.1”, i.e., about the volume of
three basketballs.
Fig. 7.7.2 shows that the molar volume of real gases is very close to that of the ideal gas. The small differences between the molar
volume of real gases and the ideal gas are because ideal gas molecules are assumed to have negligible volume and negligible
intermolecular interactions. The real gas molecules do have some volume and some intermolecular interactions that cause
deviations of real gases from the ideal behavior. However, for practical purposes, the calculations based on ideal gas law remain
applicable for the majority of real gases under ambient conditions.
7.7.3 https://chem.libretexts.org/@go/page/371605
Figure 7.7.2 : Comparison of molar volume of some real gases with an ideal gas at STP.
The molar volume of gases at STP is an equality between the number of moles and the volume of gas at STP, i.e.:
The conversion factors are used to convert volume to moles and mol to volume of gas, respectively, at STP.
Example 7.7.5
1 mol oxygen
22.41 L
volume of oxygen = 64.0 g oxygen × × = 44.82 L oxygen
32 g of oxygen 1 mol oxygen
7.7: Ideal gas law is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
7.7.4 https://chem.libretexts.org/@go/page/371605
7.8: Dalton’s law of partial pressure
Calculations in examples 2 and 3 of section 7.7 prove that 1.50 mol He in 5.1 L chamber exerts a pressure of 7.2 atm; 0.6 mol He in
the same chamber exerts a pressure of 2.9 atm, and a mixture of the two in the same chamber at the same temperature exerts a
pressure equal to the sum of the pressures that each fraction exerts if it is alone in the chamber, i.e.,
Ptotal = 7.2atm + 2.9atm = 10.1atm . What if one of the gas was hydrogen, and the other helium? The answer is: that the
calculations using the ideal gas law remain the same because it is the number of molecules, not the type of molecules that are
involved in the calculations. Properties of gases depend on the number of moles of gas n , and not on the nature of the gas, as
illustrated in Fig. 7.8.1.
Figure 7.8.1 : Illustration of Dalton’s law of partial pressures: the total pressure of a H2+He mixture equals the sum of the pressure
that of H2 and He exert alone the same size chamber at the same temperature. Source: Dr. Blair Jesse Ellyn Reich / CC BY-SA
(https://creativecommons.org/licenses/by-sa/3.0)
Atmospheric air is a mixture of nitrogen, oxygen, argon, carbon dioxide, water vapors, and trance amount of some other gases. The
atmospheric pressure is the sum of the partial pressures of components of air, as illustrated in Fig. 7.8.2.
Figure 7.8.2 : Partial pressures of main atmospheric constituents at sea level. Source: Andrew Jarvis / CC BY-SA
(https://creativecommons.org/licenses/by-sa/4.0).
Example 7.8.1
A 46 L He and 12 L O2 sample, both at 1.0 atm and 25 oC were pumped into a 5.0 L scuba diving tank at 25 oC. Calculate the
partial pressure of each gas and the total pressure in the tank?
Solution
For He before mixing: V = 46 L, T = 25 oC + 273 = 298 K, P = 1.0 atm, n He =?, R = 0.08206 L−atm
mol−K
mol−K
7.8.1 https://chem.libretexts.org/@go/page/371607
Formula and calculations: n H2 = PV/RT =
1.0 atm×12 L
L−at m
= 0.49 mol
0.08206 ×298 K
m ol−K
After mixing: V =5 L, T =25 oC+273 = 298 K, n He =1.9 mol, n O2 =0.49 mol, R = 0.08206 L−atm
mol−K
.
Requried: P = ?, P =?, and P
He O2 =? total
L −a tm L−at m
1.9 mol×0.08206 ×298 K 0.49 mol×0.08206 ×298 K
nH e RT nox y gen RT
Formula: P He =
V
= 5.0 L
m ol−K
= 9.3 atm ,PO2 =
V
= 5.0 L
m ol−K
= 2.4 atm ,
The hyperbaric chamber is an air chamber that is at two to three atmospheric pressure, as shown in Fig. 7.8.3. The solubility of
gases increases with an increase in pressure. A patient placed in a hyperbaric chamber has a higher concentration of oxygen
dissolved in blood because the partial pressure of oxygen is two to three times higher than in the atmospheric air. The higher
concentration of oxygen is toxic to many strains of bacteria. Therefore the hyperbaric chambers are used to treat burn patients,
in surgeries, and to treat some cancers.
The hyperbaric chambers are also used to treat carbon monoxide (CO) poisoning because the higher concentration of oxygen in
the chamber can displace the CO bound with hemoglobin faster than atmospheric oxygen does. Another use of hyperbaric
chambers is to treat scuba divers suffering from the bends. If a diver ascends too quickly, the nitrogen dissolved in blood makes
bubbles in the vessels that block the blood flow –a condition called bends. The divers suffering from the bends are placed in a
hyperbaric chamber at high pressure, and then the pressure is slowly decreased to atmospheric pressure. The nitrogen dissolves
in blood under higher pressure and slowly diffuses out through the lungs as the pressure is gradually decreased.
Figure 7.8.3 : The hyperbaric chamber at the Neutral Buoyancy Lab. Source: Mike / CC BY
(https://creativecommons.org/licenses/by/2.0)
7.8: Dalton’s law of partial pressure is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif
Malik.
7.8.2 https://chem.libretexts.org/@go/page/371607
CHAPTER OVERVIEW
8: Nuclear chemistry
8.1: Introduction to nuclear chemistry
8.2: Radioactivity
8.3: Half-life of radioisotopes
8.4: Radiation measurements
8.5: Ionizing radiation exposures
8.6: Medical uses of radioisotopes
8.7: Making radioisotopes for medical uses
8.8: Nuclear fusion and fission
8: Nuclear chemistry is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
1
8.1: Introduction to nuclear chemistry
What is a nuclear reaction?
Unlike chemical reactions that involve valence electrons, nuclear reactions involve changes in the nucleus of an atom, as shown
in Fig. 8.1.1.
Figure 8.1.1 : The proton-proton chain reaction, branch I, dominates in stars the size of the Sun or smaller. Source: Sarang, Public
domain, via Wikimedia Commons
Nuclear reaction
Nuclear reactions involve changes in the nucleus of an atom. A nuclear reaction may result in one or more of the following: i)
conversion of an atom to its isotope or an atom of another element, ii) conversion of mass into energy or vice versa, and iii)
release of nuclear radiations.
Although nuclear reactions are less numerous than chemical reactions, they are essential in many aspects, e.g., they are the source
of energy in the sun and stars and the synthesis of elements in the universe. Nuclear reactions are becoming essential in human life
in the form of electricity production from nuclear power plants, a source of radioisotopes for medical imaging to visualize organs
and diagnose diseases, to treat tumors, and cancerous cells, as shown in Fig. 8.1.2.
Figure 8.1.2 : Spectrum of medical imaging (left) and radiation therapy of the pelvis. (right). Source: Martin Tornai, CC BY 4.0
<https://creativecommons.org/licenses/by/4.0>, via Wikimedia Commons, and Dina Wakulchik from Indianapolis, Indiana, USA /
CC BY (https://creativecommons.org/licenses/by/2.0)
8.1.1 https://chem.libretexts.org/@go/page/372948
The composition of a nucleoid is represented by the same symbol that represents the isotopes of
elements, as X , where X is the element symbol, Z is the number of protons, and A is the number of
A
Z
protons and neutron in the nucleus. For example, carbon exists as a mixture of C, and C isotopes. 12
6
13
6
The name of the element, followed by the number of nucleons separated by a hyphen, is another way of
representing a nucleoid. For example, carbon-12, carbon-13, and carbon-14 represent the carbon
nucleoids having 6 protons each, but 6, 7, and 8 neutrons, respectively. Similarly, hydrogen exists as a
mixture of H , H , and H , that can also be represented as hydrogen-1, hydrogen-2, and hydrogen-3,
1
1
2
1
3
respectively.
Nucleons
Nuclear reaction
A nuclear reaction is a process in which two nuclei, or a nucleus and an external subatomic particle, collide to produce one or
more new nuclides.
The nuclear reaction is a reaction that involves nucleoids. The reactant nucleoid, called the parent nucleoid, usually transforms
into a different nucleoid called the daughter nucleoid. The daughter nucleoid may be an isotope of the parent nucleoid, or it may
be a different element. The conversion of an isotope to another isotope of the same or a different element is a nuclear reaction that
is called transmutation or a nuclear transformation, as shown in Fig. 8.1.3.
along with emission three protons. Copy and Paste Caption here. Source: MikeRun, CC BY-SA 4.0
<https://creativecommons.org/licenses/by-sa/4.0>, via Wikimedia Commons
Nuclear radiations
Nuclear radiation or radioactivity is the particles and energy emitted by the nucleus during a nuclear reaction.
The nuclear reaction is accompanied by the emission of nuclear radiations including high-energy electromagnetic traditions called
gamma-rays ( γ-rays), subatomic particles like electrons, positrons, protons, neutrons, or a small nucleus, like He called alpha-
4
2
2+
particles ( α -particles). Nuclear radiations are ionizing radiations, i.e., they can knock off electrons from the atoms they come in
contact with.
Radioactive
The nucleoids that are capable of spontaneous disintegration, causing the emission of nuclear radiation, are called radioactive.
The process of emission of nuclear radiation by a spontaneous disintegration of radioactive nucleoids is called radioactivity, as
illustrated in Fig. 8.1.4.
8.1.2 https://chem.libretexts.org/@go/page/372948
Figure 8.1.4 : Illustration of radioactive decay of a nucleoid. Source: Pearson Scott Foresman, Public domain, via Wikimedia
Commons
The nuclear radiations include gamma-rays ( γ-rays), alpha-particles ( α -particles), beta-particle ( β-particles), neutrons (n), and
positron (β+-particles).
Gamma-rays
The gamma-rays are electromagnetic radiations that have no mass and have energy higher than that of X-rays. The symbol γ,
γ , or γ-ray represents a gamma-ray.
0
0
Alpha-particles
The alpha-particles (α -particles) are helium nuclei with two protons, two neutrons, and without electrons, i.e., 4
2
He
2+
. The α -
particles are also represented as He or Helium-4.
4
2
Beta-particles
The beta-particles (β-particles) are fast-moving electrons that have atomic number -1, charge -1, and negligible mass. The
symbol β, β , β, or e also represents a β-particle.
− 0
−1
0
−1
Positrons
Positrons are anti-particle of electron, i.e., they have the same mass but opposite charge than that of an electron. The symbol
+β, β , β , or e represents a positron.
+ 0 0
+1 +1
8.1: Introduction to nuclear chemistry is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif
Malik.
8.1.3 https://chem.libretexts.org/@go/page/372948
8.2: Radioactivity
The cause of radioactivity
Only a particular combination of protons and neutrons forms stable nucleoids; the rest are unstable nucleoids, as illustrated in Fig.
8.2.1. The observations on the stable nucleoids are the following.
Figure 8.2.1 : Graph of isotopes by type of nuclear decay. Orange and blue nuclides are unstable, with the black squares between
these regions representing stable nuclides. The unbroken line passing below many of the nuclides represents the theoretical position
on the graph of nuclides for which the proton number is the same as the neutron number. The graph shows that elements with more
than 20 protons must have more neutrons than protons, in order to be stable. Source: Table_isotopes.svg: Napy1kenobiderivative
work: Sjlegg / CC BY-SA (https://creativecommons.org/licenses/by-sa/3.0)
The nucleoids shown as black dots in Fig. 8.2.1 are naturally occurring stable nucleoids. All nucleoids shown in colored dots other
than black are unstable.
Note
The unstable nucleoids are radioactive -they spontaneously disintegrate, i.e., they tend to re-arrange the nucleon composition in
the nucleus to become a more stable nucleoid.
The radioactivity releases energy and particles, i.e., nuclear radiation along with the re-arrangement of the nucleus, as explained in
the following sections.
8.2.1 https://chem.libretexts.org/@go/page/372949
Alpha-decay
The α -emission is one of the important processes for stabilizing heavy unstable nucleoids shown with yellow dots in Fig. 8.2.1.
Since the α -particle has two protons and two neutrons, the daughter nucleoid produced along with the α -decay has two fewer
protons and four fewer mass numbers in it than the parent nucleoid, as illustrated in Fig. 8.2.2.
Note that the γ-rays have zero mass, so they do not change the atomic number and mass number of the parent nucleoid.
Smoke detectors used in homes need α -particles for their function. Americium-241 is the α -decay-emitter used in the smoke
detectors.
Example 8.2.1
Step 2. Balance the mass number on the two sides of the equation, i.e., the mass number of unknown nucleoid is 241-4 = 237:
241 237 4
95
Am ⟶ ?
? + 2 He
Step 3. Balance the atomic number on the two sides of the equation, i.e., the atomic number of the unknown nucleoid is 95-2 =
93:
241 237 4
95
Am ⟶ 93
? + 2 He
Step 4. Find the symbol of the unknown nucleoid from the periodic table of elements, i.e., the element at atomic number 93 is
neptunium symbol Np:
8.2.2 https://chem.libretexts.org/@go/page/372949
241 237 4
95
Am ⟶ 93
Np + 2 He
This is the balanced nuclear equation for the α -decay of americium-241 in smoke detectors.
Radium-226, present in many types of rocks and soils, is an α -emitter producing radon-226 in the process. The radon-226 is
also an α -emitter that can diffuse into houses from the rocks and soil underneath the buildings. Radon is an environmental
health issue in the buildings when its concentration becomes above a certain level. The nuclear equation for the α -decay of
radon-222 is the following.
222 218 4
86
Rn ⟶ 84
Po + 2 He
Example 8.2.2
Step 2. Balance the mass number on the two sides of the equation, i.e., is 226-4 = 222:
226 222 4
88
Rn ⟶ ?
? + 2 He
Step 3. Balance the atomic number on the two sides of the equation, i.e., 88-2 = 86:
226 222 4
88
Rn ⟶ 86 ?
? + 2 He
Step 4. Find the symbol of the unknown nucleoid from the periodic table of elements, i.e., the element at atomic number 86 is
radon symbol Rn:
226 222 4
88
Rn ⟶ Rn + 2 He
?
This is the balanced nuclear equation for the α -decay of radium-226 in rocks and soil.
Beta-decay
The nucleoids marked blue in Fig. 8.2.1. have more neutrons than needed for stability. They usually stabilize them by converting
one of the neutrons (n) into a proton (p) and an electron (e) by the following nuclear process:
1 1 0
n ⟶ p+ e
0 1 −1
The proton stays in the nucleus, but the electron emits from the nucleus, as illustrated in Fig. 8.2.3. The emitted electron is called
β-particle. The process of the β-particle emission is called beta-decay. Note that the neutron has zero atomic number as there is no
proton in it, and the electron has a -1 atomic number to balance the +1 atomic number of the proton. Proton has +1 and electron has
-1 charge, which is also balanced. The mass number of an electron is zero as it has negligible mass compared to the mass of a
proton or a neutron. The electrical charges and the emission of another particle called neutrino are ignored in this equation.
8.2.3 https://chem.libretexts.org/@go/page/372949
Figure 8.2.3 : Illustration of beta-decay. A neutron (n) is transformed into a proton (p) in the nucleus, ejecting an electron (e-) and
some energy as radiation. Source: Inductiveload/ Public domain
An example of β-decay is the transformation of nitrogen-16 to oxygen-17:
16 16 0
7
N ⟶ 8
O+ e
−1
Note that in the β-decay process, the mass number remains the same, but the atomic number increases by one in the daughter
nucleus. The nuclear equation is balanced because the mass number is the same (16 = 16+0), and the atomic number is also the
same (7 = 8-1) on the two sides of the equation.
Example 8.2.3
Write the nuclear equation for the β-decay of iodine-131.
Solution
Step 1. The symbol and the atomic number of iodine in the periodic table are I and 53, respectively. So the initial equation is:
131 ? 0
53
I ⟶ ?
?+ e
−1
Step 2. Balance the mass number on the two sides of the equation, i.e., the mass number of
unknown nucleoid is 131-0 = 131:
131 131 0
I ⟶ ?+ e
53 ? −1
Step 3. Balance the atomic number on the two sides of the equation, i.e., the atomic number of
the unknown nucleoid is 53-(-1) = 54:
131 131 0
53
I ⟶ 54
?+ e
−1
Step 4. Find the symbol of the unknown nucleoid from the periodic table of elements, i.e., the element at atomic number 54 is
xenon symbol Xe:
131 131 0
53
I ⟶ 54
Xe + e
−1
This is the balanced nuclear equation for the α-decay of iodine-131 that is used to treat over-active thyroid glands.
8.2.4 https://chem.libretexts.org/@go/page/372949
Example 8.2.4
Step 2. Balance the mass number on the two sides of the equation, i.e., the mass number of unknown nucleoid is 90-0 = 90:
90 90 0
Y ⟶ ?+ e
39 ? −1
Step 3. Balance the atomic number on the two sides of the equation, i.e., the atomic number of the unknown nucleoid is 39-(-1)
= 40:
90 90 0
39
Y ⟶ 40
?+ e
−1
Step 4. Find the symbol of the unknown nucleoid from the periodic table of elements, i.e., the element at atomic number 40 is
zirconium symbol Z:
90 90 0
39
Y ⟶ 40
Z+ e
−1
Positron emission
The nucleoids marked orange in Fig. 8.2.1. have more protons than needed for stability. They usually stabilize them by converting
one of the protons (p) into a neutron (n) and a positron β+ by the following nuclear process:
1 1 0
p ⟶ n+ e
1 0 1
The neutron stays in the nucleus, but the positron emits from the nucleus, as illustrated in Fig. 8.2.4. Note that the positron has a +1
mass number that balances the +1 atomic number of the proton on the other side of the equation. The positron has a +1 charge that
also balances the +1 charge of the proton on the other side of the equation. The mass number of a positron is zero as it has
negligible mass compared to the mass of a proton or a neutron. The electrical charges are not shown in the nuclear equation.
Figure 8.2.4 : Illustration of positron-emission. A proton (p) is transformed into a neutron (n) in the nucleus, ejecting a positron (e+)
and some energy as radiation. Source: Master-m1000 / Public domain
Carbon-11 is an example of positron-emitter:
11 11 0
6
I ⟶ 5
B+ e
1
Note that in the positron-emission process, the mass number remains the same, but the atomic number decreases by one in the
daughter nucleus. The nuclear equation is balanced because the mass number is the same (11 = 11+0), and the atomic number
is also the same (7 = 5+1) on the two sides of the equation.
8.2.5 https://chem.libretexts.org/@go/page/372949
Uses of some positron emittors
Positron emission is used in positron emission tomography (PET) which is a medical imaging technique. Short-lived positron-
emitting isotopes 11C, 13N, 15O, and 18F used for positron emission tomography are typically produced by proton irradiation of
natural or enriched targets described in a later section.
Fluorine-18 in fluorodeoxyglucose, abbreviated as [18F]FDG is a positron emitter commonly used to detect cancer, and in
[18F]NaF is widely used for detecting bone formation. Other examples are oxygen-15 in [15O]H2O used to measure blood flow
and nitrogen-13 used to tag ammonia molecules for myocardial perfusion imaging.
15 15 0
O ⟶ N+ e
8 7 1
13 13 0
N ⟶ 6
C+ e
7 1
Example 8.2.5
Step 2. Balance the mass number on the two sides of the equation, i.e., the mass number of
unknown nucleoid is 18-0 = 131:
18 131 0
9
F ⟶ ?
?+ e
1
Step 3. Balance the atomic number on the two sides of the equation, i.e., the atomic number of the unknown nucleoid is 9-(+1) = 8:
18 131 0
F ⟶ ?+ e
9 8 1
Step 4. Find the symbol of the unknown nucleoid from the periodic table of elements, i.e., the element at atomic number 8 is
oxygen symbol O:
18 18 0
9
F ⟶ 8
?+ e
1
Gamma-emission
The gamma-rays are high-energy electromagnetic radiations that do not have mass or charge. So, pure γ-emission happens from the
nucleus, but it does not result in transmutation, simply the nucleoid changes from a more unstable state, called a metastable state,
to a relatively stable state, as illustrated in Fig. 8.2.5.
Figure 8.2.5 : An illustration of gamma-emission, the γ -ray originates from the radioactive nucleus, changes the energy state but
does not change the identity of the nucleoid. Source: Inductiveload / Public domain
A symbol m or * next to the mass number as a superscript to the right indicates the metastable state of the parent nucleoid. For
example, technetium-99m is a γ-emitter widely used in medical imaging:
8.2.6 https://chem.libretexts.org/@go/page/372949
99 m 199
Tc ⟶ Tc + γ
43 43
Note that the nucleoid remains the same after γ-emission, except for the change form metastable to a more stable state indicated by
m.
Often, the γ-emission accompanies α -emission or β-emission. For example, polonium-210 decays by a simultaneous α -emission
and γ-emissions.
210 206 4
84
Po ⟶ 82
Pb + 2 He + γ
Similarly, iridium-192 used in implants to treat breast cancer, and cobalt-60 used as an external radiation source for cancer
treatment, simultaneously emit β and γ-rays.
192 192 0
Ir ⟶ Pt + e+γ
77 78 −1
60 60 0
Co ⟶ Ni + e+γ
27 28 −1
Iodine-131 decays to β-particle and xenon-131m that is rapidly followed by a γ-decay of xenon-131m.
131 131 m 0
I ⟶ Xe + e
53 54 −1
131 m 131
Xe ⟶ Xe + γ
54 54
Note that the mass number remains the same, but the atomic number decreases by one in the electron-capture process.
Chromium-51, which is used for imaging the spleen, decays by the electron capture and γ-emission.
51 0 51
Cr + e ⟶ V+γ
24 −1 23
8.2.7 https://chem.libretexts.org/@go/page/372949
Figure 8.2.6 : Transition diagram for decay modes of a radionuclide, with neutron number N and atomic number Z (shown are α , ,
β± are beta-particle, and positron,p+is proton, and n0 is neutron missions, EC denotes electron capture). Source: MarsRover / GFDL
(http://www.gnu.org/copyleft/fdl.html)
8.2: Radioactivity is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
8.2.8 https://chem.libretexts.org/@go/page/372949
8.3: Half-life of radioisotopes
The half-life (t1/2) of a radioisotope is the time it takes for half of the sample to decay.
It tells the rate of decay of the radioisotope – the faster the rate of decay, the shorter the half-life.
Figure 8.3.1 : Graph of the stability of every known nucleoid. Plotted as Z (number of protons) versus N (number of neutrons). The
color corresponds to the value of the half-life t½ with a strong log scale, as it varies between 10-20 and 1020 seconds. Source:
Fffred~commonswiki/ Public domain
Table 1: Half-lives (t1/2) of some common radioisotopes
Radioisotope Symbol Half-life Use
Carbon-14 14
6
C 5730 years Radioisotope dating
Hydrogen-3 3
1
H 12.3 years Radioisotope dating
8.3.1 https://chem.libretexts.org/@go/page/372950
Radioisotope Symbol Half-life Use
Potassium-40 40
19
K 1.3 x 109 years Radioisotope dating
Rhenium-187 187
75
Re 4.3 x 1010 years Radioisotope dating
Uranium-238 238
92
U 4.5 x 109 years Radioisotope dating
Uranium-235 235
92
U 7.0 x 108 years Nuclear reactor fuel
Cobalt-60 60
27
Co 5.3 years Medical (external radiation source)
Iodine-131 131
53
I 8.1 days Medical
Iron-59 59
26
Fe 45 days Medical
Molybdenum-99 99
42
Mo 67 hours Medical
Sodium-24 24
11
Na 15 hours Medical
Technetium-99m 99 m
43
Te 6 hours Medical
Phosphoros-32 32
15
P 14.3 dsys Medical
Suppose there is 100 mg of the phosphorous-32 in the beginning; 50 mg will be left behind after 14.3 days, i.e., after 1 half-life;
and 25 mg will be left after 28.6 days, i.e., after 2 half-lives. A negligible amount of the parent isotope phosphorous-32 is left after
9 half-lives.
Figure 8.3.2 : Decay of phosphorous-32 (blue color) and accumulation of sulfur-32 (orange color) after each half-life.
The amount of a radioisotope remaining after the given time can be calculated from the known initial amount and time spent, by
the following formula:
n
mf = mi (0.5 )
where mi is the initial amount, mf is the final amount, and n is the number of half-lives passed. The formula works even if the
number of half-lives is not a whole number.
8.3.2 https://chem.libretexts.org/@go/page/372950
Example 8.3.1
If 50.0 mg of iodine-131 was injected for medical treatment, how many milligrams will be left after 40.5 days? (Half-life of
iodine-131 is 8.1 days)
Solution
Given: mi = 50.0 mg, Time = 40.5 days, Desired ? mf
The equality: 1 half-live = 8.1 days, gives the following conversion factors.
1 half-life 8.1 days
and
8.1 days 1 half-life
For calculating the half-lives, multiple the given time with the conversion factor that cancels
the time:
1 half-life
n = 40.5 days × = 5 half-lives
8.1 days
For calculating the amount left, plug in the values in the formula:
n 5
mf = mi (0.5 ) = 50.0mg(0.5 ) = 1.56 mg
Radioisotope dating
Natural radioactivity is used to establish the age of objects of archeological, anthropological, or historic interest. All living objects
have carbon in their composition. Carbon-14 is a radioactive isotope of carbon with a half-life of 5730 years. Carbon-14 is
produced by the transmutation of nitrogen-14 upon neutron bombardment by cosmic rays, as illustrated in Fig. 8.3.3. Its
concentration in a carbon source for the living organism remains almost constant because its decay counterbalances its production
by cosmic rays. Living organisms continuously replenish carbon, so the carbon-14 concentration remains almost constant as long
as the object is alive. After the object dies, the carbon-14 decreases with time, reducing to half after one half-life. The carbon-12
isotope is not radioactive, so its concentration remains constant. Measurement of the carbon-14/carbon-12 ratio allows calculating
the age of the object after its death. The age of early civilizations, like the Indus valley civilization examples shown in Fig. 8.3.4
were determined by the carbon-14 dating method.
8.3.3 https://chem.libretexts.org/@go/page/372950
Figure 8.3.4 : Excavated ruins of Mohenjo-daro, Sind province of Pakistan, showing the Great Bath in the foreground. Mohenjo-
daro, on the right bank of the Indus River, is a UNESCO World Heritage Site. Dated 3300 BCE to 1300 BCE. Source: Saqib
Qayyum / CC BY-SA (https://creativecommons.org/licenses/by-sa/3.0)
Because the carbon-14 decreases with time, the object older than 20,000 years of age does not have sufficient carbon-14 left to
determine their age accurately. Other radioisotopes with a longer half-life, e.g., uranium-238 with a half-life of 4.5 x 109 years, are
used to determine the age of ancient objects. For example, the age of rock samples from the moon, as shown in Fig. 8.3.5 was
determined by uranium-238 radioisotope dating.
Figure 8.3.5 : Lunar Olivine Basalt sample collected from the moon by the Apollo 15 mission, at station 9A on the rim of Hadley
Rille. It was formed around 3.3 billion years ago. On display in the National Museum of Natural History. Source: Wknight94 talk /
CC BY-SA (https://creativecommons.org/licenses/by-sa/3.0)
8.3: Half-life of radioisotopes is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
8.3.4 https://chem.libretexts.org/@go/page/372950
8.4: Radiation measurements
Radioactivity measurement
Radioactivity is measured in terms of the rate of radioactive events. Nuclear radiations are ionizing radiations, i.e., they knock off
electrons from atoms or molecules that come in their path, leaving behind cations. Geiger Muller counter is one of the radiations
measuring instruments that counts the disintegration of radionucleotide per second by registering the current produced by the
ionization action of the radiation, as illustrated in Fig. 8.4.1. It is not just one ionization event; the nuclear particle keeps ionizing
the atoms in its track until its energy is exhausted, as illustrated in Fig. 8.4.2. The instrument records the flash of electric current
produced by the ionization of each radioactive disintegration.
Figure 8.4.2 : Tracks of α-particles from americium-241 made visible in a cloud chamber. Source: https://youtu.be/noP7HT-Uins
Becquerel (Bq)
The SI unit of radioactivity is Becquerel (Bq), i.e., the number of nuclei that disintegrate per second.
The common unit of radiation intensity is Curie (Ci), i.e., 3.7 x 1010 disintegrations per second. The relationship between Becquerel
and Curie is the following.
10 10
1Ci = 3.7 × 10 Bq = 3.7 × 10 disintegrations
Often the radioisotope for medical use has the information of millicurie per milliliter (mCi/mL) from which the volume for the
desired dose can be calculated.
Example 8.4.1
A patient must be given a 5.0 mCi dose of iodine-131 that is available as Na131I solution containing 3.8 mCi/mL. What volume
of the solution should be administered?
Solution
Use the reciprocal of 3.8 mCi/mL as a conversion factor:
8.4.1 https://chem.libretexts.org/@go/page/372951
1 mL
5.0 mCi × = 1.3 mL dose
3.8 mCi
Gray (Gy)
The SI unit of absorbed dose is gray (Gy) which is defined as the absorption of one joule of radiation energy per kilogram of
matter (J/kg).
The common unit of absorbed dose is rad, which stands for radiation absorbed dose. The rad is one-hundredth of a gray, i.e.:
1 Gy = 100 rad
Equivalent dose
The same amount of energy deposited in tissues by different types of radiation carries different levels of health risks in terms of
causing cancer and genetic damage, expressed as a radiation weighting factor (WR), as illustrated in Fig. 8.4.4, and listed in Fig.
8.4.5. For example, 1 Gy of beta -particles carries a risk of 5.5% chances of eventually developing cancer, while 1 Gy of alpha-
particles has 20 times more risk compared to the β-particle (ref.: https://en.Wikipedia.org/wiki/Sievert, accessed on 07/15/2020).
The health risk of the ionizing radiation is measured in the units of equivalent dose. Sievert (Sv) is an SI unit of an equivalent dose
of ionizing radiation that measures the health effects of low levels of ionizing radiation on the human body.
Sievert (Sv)
The equivalent doze in Sievert (Sv) is equal to the product of absorbed dose in grays (Gy) multiplied the radiation weighting
factor (WR), i.e., The equivalent dose in Sv = Absorbed dose in Gy × W R
8.4.2 https://chem.libretexts.org/@go/page/372951
Figure 8.4.4 : Illustration of radiation equivalent dose, effective dose, radiation weight factors (WR), and tissue weighting factors
(WT). Source: Ministry of the Environment of Japan website, https://www.env.go.jp/en/chemi/rhm/b...g-02-03-04.png
Figure 8.4.5 : Radiation weighting factors WR used to represent relative biological effectiveness according to ICRP report 103,
Source of data: https://en.Wikipedia.org/wiki/Sievert, accessed on 07/16/2020
The common unit of equivalent dose is rem (rem stands for roentgen equivalent man), which is:
1 Sv = 100rem
The personnel working in a radiation environment are required to wear film badges or electronic personal dosimeters, as shown in
Fig. 8.4.6, that record the dose received. A record of each person's dose is usually maintained by the radiation facilities to comply
with the allowed radiation exposure limits.
Figure 8.4.6 : View of readout on an electronic personal dosimeter. The clip is used to attach it to the wearer's clothing. Source:
Rama / CC BY-SA 2.0 FR (https://creativecommons.org/licenses...2.0/fr/deed.en)
Effective dose
The equivalent dose is equal to the effective dose in sievert (Sv) when the whole human body is exposed equally to the radiation. If
part of the body is exposed, then an effective dose in sievert (Sv) is calculated by the summation of the product of equivalent dose
8.4.3 https://chem.libretexts.org/@go/page/372951
in Sv with tissue weighting factor (WT) for each tissue exposed to the radiation, as illustrated in Fig. 8.4.4 and calculated with
example in Fig. 8.4.7. The reason for this calculation is that the effect of the same equivalent dose is different in different tissues.
The tissue weighting factors (WT) are listed in Table 1.
An effective dose takes the absorbed dose and adjusts it for radiation type and organ sensitivity, i.e.,:
Figure 8.4.7 : Example calculations of effective dose. Source: Ministry of the Environment of Japan website,
https://www.env.go.jp/en/chemi/rhm/b...g-02-03-06.png
Table 1: Tissue weighting factors (WT) ICRP103 (2007), source of data: Source: https://en.Wikipedia.org/wiki/Sievert, accessed on
07/16/2020
Organs WT
Gonads 0.08
Colon 0.12
Lung 0.12
Stomach 0.12
Breasts 0.12
Bladder 0.04
Liver 0.04
Oesophagus 0.04
Thyroid 0.04
Skin 0.01
Brain 0.01
Total 1.00
8.4: Radiation measurements is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
8.4.4 https://chem.libretexts.org/@go/page/372951
8.5: Ionizing radiation exposures
Interaction of radiation with matter
The ionizing radiations knock off electrons from atoms and molecules. Water composes about 60% of the human body. The oxygen
atom in water has eight valence electrons, four in the two bonding pairs, and four in the two lone pairs, as shown in Fig. 8.5.1. If
one of them is knocked off by radiation, the result is a radical-cation. It is radical because it has one unpair electron and its octet is
incomplete, and cation because an electron has been lost. The radical cation can then release a proton to become a hydroxyl radical.
The radicals are very reactive species due to the incomplete octet. The radicals tend to react with any material around that causes
damage to the tissues. Notably, the damage to DNA is the most dangerous, causing mutation, cancer, and hereditary problems.
Figure 8.5.1 : Ionization radiations knock off an electron from molecules like water, producing very reactive radical cations that
lead to the production of very reactive radicals like hydroxyl radical in this case.
The cells that are proliferating are more susceptible to the harmful effects of radiation exposure, including bone marrow, skin,
reproductive organs, and intestinal lining, as well as all the cells of growing children. If the bone marrow cells are damaged, the red
blood cells may not be produced. Damage to the reproductive cells or the cells of a fetus may cause congenital disabilities.
Fortunately, the cancer cells are multiplying and affected by the radiation exposure much more than the surrounding healthy cells,
which allows for selectively killing the cancerous cells by radiation treatment. The use of ionizing radiations in cancer treatment is
described in a later section.
8.5.1 https://chem.libretexts.org/@go/page/372953
Figure 8.5.2 Source: National Environment Agency, Open
: Health Effects of Ionising Radiation on People.
Data Licence, https://www.nea.gov.sg/our-services/...tion-on-people
Background radiation exposures
Humans are exposed to radiation in the cases of nuclear accidents, during nuclear radiation treatments, particularly for treating
different forms of cancers, and during medical imaging for medical diagnostic purposes. Besides these human-made radiation
exposures, humans are regularly exposed to natural sources of radiation, called background radiation. The background radiation
can be in foods, e.g., potassium-40 is a naturally occurring radioactive isotope present in potassium-containing foods. Carbon-14,
radon-222, strontium-90, and iodine-131 are other radioisotopes present in the air and foods around us. Radioactive isotopes of
uranium and thorium and their decay products are the source of radiation in soil. The extraterrestrial sources of radiation are cosmic
rays that are stopped in the upper atmosphere, but some may reach the ground and expose the people. The average annual radiation
dose per person in the U.S. is 6.2 millisieverts (620 millirem).
8.5.2 https://chem.libretexts.org/@go/page/372953
Figure 8.5.3 : Source: Ministry of the Environment of Japan website, https://www.env.go.jp/en/chemi/rhm/b...g-01-03-08.png
Radiation protection
Different types of radiation have different penetration depths in the air and the human body. The α -particles are heavy with two
protons and two neutrons; they cause much ionization but travel a few centimeters in air. A sheet of paper can stop the α -particles,
as illustrated in Fig. 8.5.3. Exposure to α -particles from an external source affects only the outer layer of the skin. However, an
internal source can cause significant damage to the nearby tissues, as illustrated in Fig. 8.5.4
8.5.3 https://chem.libretexts.org/@go/page/372953
The workers in a radiation environment wear heavy clothes, gloves, and lab coats to provide additional protection. The radioactive
materials are usually stored in shielded containers, even the syringes containing radioactive materials for injection are shielded. The
general rules for protection against the radiation are:
8.5: Ionizing radiation exposures is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
8.5.4 https://chem.libretexts.org/@go/page/372953
8.6: Medical uses of radioisotopes
The use of nuclear chemistry in medical technologies is increasing over time. The medical uses can be divided into two categories :
1. medical imaging of organs or diagnosing any malfunction and
2. therapeutic use, mainly for killing cancerous cells.
Table 8.6.1 lists some radioisotopes commonly used in medical imaging. A low dose of the radioisotope is administered to a
patient. The γ-rays cross over the body and are recorded like X-rays. A computer finally converts the recording into a useful image.
The image is compared with an image of a healthy organ to diagnose any malfunction.
Table 8.6.1: Some of the commonly used radioisotopes in medical imaging
Radioisotope Symbol Mode of decay Half-life Use in medical imaging
Carbon-11 11
6
C β+, g 20.3 m Brain scan to trace glucose
Fluorine-19 18
9
F β+, g 109m Brain scan to trace glucose
Selenium-75 75
34
Se E.C., g 118 d Pancreas scan
Krypton-81m 81 m
36
Kr g 13.3 s Lung ventilation scan
Xenon-133 133
54
Xe β 5.24 d Lung ventilation scan
Mercury-197 197
80
Hg E.C., g 64.1 h Kidney scan
Phosphorous-32 32
15
P β 14.3 d Detect eye tumors
An example of medical imaging is the thyroid gland in the neck that produces the hormone thyroxin, which controls the overall rate
of metabolism in the body. Each thyroxin molecule contains four iodine atoms. Administration of radioactive Na131I or Na123I salt
accumulates the iodine in the thyroid gland in a few hours. Decay of 131I and 123I involves γ-emission.
131 131 0
53
I ⟶ 54
Xe + e+γ
−1
8.6.1 https://chem.libretexts.org/@go/page/372954
123 0 123
I+ e ⟶ Xe + γ
53 −1 52
The γ-emission from the iodine localized in the thyroid gland is recorded, as shown in Fig. 8.6.1. An overactive thyroid
(hyperthyroidism) cumulates more and underactive thyroid (hypothyroidism) cumulates less iodine than a healthy thyroid.
Figure 8.6.1 : Thyroid scan with iodine-123 for evaluation of hyperthyroidism. Source: Myohan at en.Wikipedia / CC BY
(https://creativecommons.org/licenses/by/3.0)
Another example is positron emission tomography (PET). Positron emitters like carbon-11 and fluorine-18 incorporated in a
suitable compound like glucose allow following the metabolic path of the compound. For example, 18-fluorodeoxyglucose )18-
FDG) is a glucose molecule in which one of the oxygen is replaced with 18F. Intravenous injection of the 18-FDG ultimately results
in the cumulation of 18-FDG in the brain and other body organs where glucose is used in the metabolic process. The 18F emits a
positron, which, being an anti-particle of the electron, reacts with the electron and produces two g γ-rays.
18 18 0
9
F ⟶ 8
O+ e+γ
1
0 0
e+ e ⟶ 2γ
1 1
The γ-rays are used to obtain an image of the organ. The image reveals problem areas in the form of an abnormal concentration of
glucose in the part of the organ. For example, Fig. 8.6.2 compares the PET image of a healthy brain versus a brain with Alzheimer's
disease.
Figure 8.6.2 : PET scan of bran: healthy (left) and Alzheimer's disease (right). Source: US National Institute on Aging,
Alzheimer's Disease Education and Referral Center / Public domain
Since glucose metabolism happens in all organs, whole-body PET scans can be used to diagnose lung, colorectal, head and neck,
and esophageal cancers as well as other diseases that involve abnormal glucose metabolism. For example, tumors have high
metabolic rates; the PET scans using 18-FDG are used to detect them, as shown in Fig. 8.6.3.
File:PET-MIPS-anim.gif
Figure 8.6.3 : Whole-body PET scan using 18F-FDG. The normal brain and kidneys are labeled, and radioactive urine from
breakdown of the FDG is seen in the bladder. In addition, a large metastatic tumor mass from colon cancer is seen in the liver.
Source: Jens Maus (http://jens-maus.de/) / Public domain
8.6.2 https://chem.libretexts.org/@go/page/372954
3. Magnetic resonance image (MRI) is another powerful medical imaging technique that is based on the fact that hydrogen atoms
splint into two two energy states when placed in a strong magnetic field. When illuminated with infrared (IR) radiation of the
energy matching with the energy gap between the two groups, the hydrogen atoms are excited from the lower to higher energy
state. The decay of the excited state emits the IR radiations that are recorded to obtain the image of soft tissues that contain
many hydrogen atoms in the form of water molecules. Fig. 8.6.4 shows some examples.
Figure 8.6.4 : Spectrum of medical imaging. Source: Martin Tornai, CC BY 4.0 <https://creativecommons.org/licenses/by/4.0>, via
Wikimedia Commons
Radiation therapy
The purpose of radiation therapy is to selective kill the diseased cells or tissues by exposing them to radiation. Higher radiation
doses are required for therapy than for imaging. The radiation source can be external or internal.
Figure 8.6.5 : Radiation therapy for Hodgkin's Lymphoma in a Versa HD.. Souce: Jakembradford, CC BY-SA 4.0
<https://creativecommons.org/licenses/by-sa/4.0>, via Wikimedia Commons
8.6.3 https://chem.libretexts.org/@go/page/372954
Figure 8.6.6 : Body sites in which brachytherapy can be used to treat cancer. Source: Rock mc1 / CC BY-SA
(https://creativecommons.org/licenses/by-sa/3.0)
Brachytherapy
Brachytherapy or seed implantation is a form of internal radiation treatment that delivers a high dose for a short period compared to
the external radiation treatment. Fig. 8.6.6 shows the sites in the body where brachytherapy can be used to treat cancer. For
example, 40 or more titanium capsules, about the size of a rice grain (Fig. 8.6.7), are implanted to treat prostate cancer. The seed
contains γ-emitter like iodine-125 (half-life 60 days), palladium-103 (half-life 17 days), or cesium-131 (half-life 10 days). The seed
may be left in the body because, due to the short half-life, they are no more significantly radioactive after the treatment.
Figure 8.6.7 : Photo of radioactive seeds. These implants are a form of radiation therapy for prostate cancer. Source: Nuclear
Regulatory Commission from US / Public domain
Another option is temporary brachytherapy, e.g., iridium-192 needles that deliver higher radiation doses are inserted to treat
prostate cancer and removed after 5 to 10 min. The iridium-192 wires are also used as a follow-up treatment after breast cancer
surgery to kill any residual or recurring cancer cells. The iridium-192 wires are inserted through a catheter implanted in the space
from where the tumor was removed. The wires are removed after delivering the required radiation dose. The process is repeated
twice a day for five days. The catheter is removed, and no radioactive material is left in the body after the treatment.
8.6: Medical uses of radioisotopes is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
8.6.4 https://chem.libretexts.org/@go/page/372954
8.7: Making radioisotopes for medical uses
Natural radioisotopes usually have a long half-life and are not best suited for medical applications. The medical application usually
requires short-lived radioisotopes. The radioisotopes are usually produced in nuclear reactors where particles, like α -particles, β-
particles, and neutrons, are abundant. Particle accelerators, such as the one shown in Fig. 8.7.1 also accelerate and direct the
nuclear particles at the targets. The high-energy nuclear particles may be absorbed by and transmute the target nuclei to
radioisotopes in a nuclear reaction.
Figure 8.7.1 : SLAC's particle accelerator may be two miles long, but researchers at FACET are working to develop more compact
versions that could be widely used in medicine and industry -- particle accelerators are used for cancer research, processing
computer chips, and even producing the shrink wrap used to keep your Thanksgiving turkey fresh. Source: ENERGY.GOV / Public
domain
Radioisotopes in medical applications are usually produced by the particle bombardment method. One reaction that happens
naturally by neutron bombardment from cosmic rays on nitrogen-14 is the following.
14 1 14 1
N+ n ⟶ C+ p
7 0 6 1
An example of an artificial nuclear reaction initiated by α -particle bombardment on nitrogen, observed by Rutherford that lead to
the discovery of proton, is illustrated in Fig. 8.7.2.
4 14 17 1
2
He + 7
N ⟶ 8
O+ p
1
Figure 8.7.2 : A nuclear reaction observed by Rutherford using a cloud chamber. When an alpha ray strikes nitrogen, reaction "a"
(proton knock-off) does not occur; reaction "b" (α-p) occurs.). Source: / Public domain
Another example is the nuclear reaction initiated by α -particles on beryllium, observed by James Chadwick, which lead to the
discovery of the neutron.
4 9 12 1
2
He + Be ⟶ 6
C+ n
4 0
An example of radioisotope production for medical uses is the following. Gold-198, used as a tracer in the liver, is produced by
neutron bombardment on gold-197.
197 1 198
79
Au + n ⟶ 79
Au
0
8.7.1 https://chem.libretexts.org/@go/page/372955
66 1 67
30
Zn + p ⟶ 31
Ga
1
Molybdenum-99 is also produced as a fission product of uranium-235. Molybdenum-99 decays to technetium-99m that has several
uses in nuclear medical imaging and treatment.
99 99 m 0
42
Mo ⟶ 43
Tc + e
−1
Figure 8.7.1 : The first technetium-99m generator developed at Brookhaven National Laboratory, circa 1958, shown without
shielding. A Tc-99m pertechnetate solution is being eluted from Mo-99 molybdate bound to a chromatographic substrate. Source:
Brookhaven National Laboratory. / Public domain
Technetium-99m is short-lived (half-life 6 h), and needs to be produced in the hospital to minimize its decay during the transport.
Its parent molybdenum-99 has a half-life of 66h and can be transported without significant decay during the transport.
Molybdenum-99/technetium-99m generators are supplied to the hospitals in a shielded container. Fig. 8.7.3 illustrates the first
Molybdenum-99/technetium-99m generator developed at Brookhaven National Laboratory. Molybdate (MoO42) ion is adsorbed
onto alumina adsorbent in a column. When molybdenum-99 decays to technetium-99m, the ion change to pertechnetate (TcO4‑),
which is less tightly bound to the alumina. Pouring a saline solution through the column elutes the technetium-99m as TcO4‑ ion,
which is then used for medical purposes in the hospitals.
Destruction of an inoperable tumor has also been tested by α -emission from boron-10 upon neutron bombardment.
1 10 7 4
n+ 5
B ⟶ Li + 2 He
0 4
8.7: Making radioisotopes for medical uses is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad
Arif Malik.
8.7.2 https://chem.libretexts.org/@go/page/372955
8.8: Nuclear fusion and fission
Conversion of matter into energy in nuclear reactions
When free nucleons come together to make a nucleoid, they release energy equal to the nuclear binding energy. The energy is
produced at the expense of the mass of the nucleons following the famous Einstein equation:
2
E = mc
, where E is the energy, m is the mass, and c is the speed of light. In other words, the mass of nucleons in nucleoids is slightly less
than the mass of the free nucleons, as some of their mass is released as nuclear binding energy.
The nuclear binding energy per nucleon is different for different nucleoids, as shown in Fig. 8.8.1. The composition of nucleoids
changes during the nuclear reaction, and the difference in the nuclear binding energy is released as energy during the process,
which is the source of energy in the sun, stars, nuclear power plants, and nuclear weapons.
Figure 8.8.1 : Graph of the binding energy released by an atom as a function of its nucleons, the energy is in megaelectron volts
(MeV) where 1 Mev = 1.60218 10-13 J. Source: Modified from Fastfission/ Public domain
The most stable nucleoid with the highest average binding energy per nucleon is iron-56.
Nuclear fusion
The lighter nucleoids tend to combine and make heavier nucleoids that are more stable and lease a tremendous amount of
energy –a process called nuclear fusions.
Nuclear fission
The heavier nucleoids, particularly those having a mass number higher than 92, tend to split into two smaller nucleoids that are
more stable and lease a large amount of energy –a process called nuclear fission.
Both the fusion and fission processes release a tremendous amount of energy, called nuclear energy or atomic energy, which is the
difference between the nuclear binding energy of the parent and the daughter nucleoid.
8.8.1 https://chem.libretexts.org/@go/page/372956
1 2 3
H+ H ⟶ H
1 1 2
3 3 4 1
2
H + 2H ⟶ 2
H+2 H
1
3 1 4 0
H+ H ⟶ H+ e
2 1 2 1
Although hydrogen is abundant on earth and a potential source of energy, the controlled fusion reaction is not yet economically
feasible. The main difficulty is that the fusion of nuclei requires extremely high temperature, greater than a hundred thousand
Celsius, and extremely high pressure, greater than a hundred thousand atmospheres, to overcome the repulsive forces of the like
charges of the nuclei before they can fuse. For this reason, nuclear fusion reactions are also called thermonuclear reactions. Such
conditions exist in sun and stars where these reactions routinely happen, but there are many technical problems yet to overcome
before the energy can be harnessed economically from the nuclear fusion reactions on earth. A nuclear explosion can create the
conditions needed for the thermonuclear reactions and are used to carry out uncontrolled thermonuclear reactions to boost the
explosive power of nuclear weapons. Such a nuclear weapon is called a thermonuclear or hydrogen bomb.
Figure 8.8.2 : Fusion of deuterium with tritium creating helium-4, freeing a neutron, and releasing 17.59 MeV as kinetic energy of
the products while a corresponding amount of mass disappears, in agreement with kinetic E = Δmc , where Δm is the decrease in
2
the total rest mass of particles. Source: Wykis (talk · contribs), Public domain, via Wikimedia Commons
One of the potential candidate reactions for nuclear fusion reactors is the fusion of deuterium with tritium, illustrated in Fig. 8.8.2,
which requires a little less harsh conditions. Most of the research is focused on the device called Tokamak, which uses strong
magnetic fields to contain and heat the materials for the reaction, but so far it is not economically feasible to use it as a source of
energy for commercial uses.
Nuclear fission -the source of energy in nuclear power plants and in nuclear weapons
Nuclear fission is a radioactive decay process in which a heavier nucleoid splits into two or lighter nucleoids and releases a
tremendous amount of energy that is the difference in the binding energy of nucleons in the daughter and the parent nucleoids. The
nucleoids that do spontaneous fission are marked green in Fig. 8.2.1.
Nuclear fuel
Induced fission is fission that occurs upon bombardment of a nucleoid with a nuclear particle, like a neutron. For example, the
bombardment of uranium-235 or plutonium-239 by a neutron causes the fission and releases about three neutrons along with the
smaller nucleoids, as illustrated in Fig. 8.8.3. The neutrons released by the fission cause fission of another nucleoid starting a
nuclear chain reaction. The nuclei like uranium-235 or plutonium-239 that produce neutrons in the product that are capable of
continuing the fission in a chain reaction are called fissile. The fissile nucleoids are nuclear fuels used in nuclear power plants to
produce nuclear energy.
8.8.2 https://chem.libretexts.org/@go/page/372956
Figure 8.8.3 : Illustration of a nuclear fission chain reaction. Source: MikeRun / CC BY-SA
(https://creativecommons.org/licenses/by-sa/4.0)
The neutrons released in a fission event may be lost, absorbed by another nucleus like uranium-238 that does not fission, or
may be absorbed by another fissile nucleoid and repeat the fission, as illustrated in Fig. 8.8.4. Three situations may arise, If, on
the average;
1. less than one neutron released causes new fission, the process dyes,
2. one neutron released causes new fission, the fission continues in a controlled way as in nuclear power plants,
3. more than one neutron released causes new fission, the fission increases exponentially resulting in a nuclear explosion, as
illustrated in Fig. 8.8.3.
Figure 8.8.4 : A schematic nuclear fission chain reaction. 1. A uranium-235 atom absorbs a neutron and fissions into two new
atoms (fission fragments), releasing three new neutrons and some binding energy. 2. One of those neutrons is absorbed by an atom
of uranium-238 and does not continue the reaction. Another neutron is simply lost and does not collide with anything, also not
continuing the reaction. However, the one neutron does collide with an atom of uranium-235, which then fissions and releases two
neutrons and some binding energy. 3. Both of those neutrons collide with uranium-235 atoms, each of which fissions and releases
between one and three neutrons, which can then continue the reaction. Source: Fastfission/ Public domain
Note that each fission event does not necessarily split the parent nucleoid into exact same daughter nucleoids, it can split into
different combinations of daughter nucleoids resulting in a mixture of products that are usually radioactive, as illustrated in Fig.
8.8.4.
8.8.3 https://chem.libretexts.org/@go/page/372956
Atomic bomb
A certain minimum mass of the fissile material, called critical mass, is necessary for a fission event to grow exponentially and
cause a nuclear explosion. The critical mass is less for a more enriched fissile material. Therefore, usually enriched uranium-235 or
plutonium 239 is used as an explosive in an atomic bomb. The mass more than the critical mass, called supercritical mass, is split
into sub-critical portions which are kept separate in the weapon to avoid a nuclear explosion. When needed, the conventional
chemical explosion is used to force the sub-critical portions to combine and make a supercritical mass for a nuclear explosion to
happen, as illustrated in Fig. 8.8.5.
Figure 8.8.5 : Schematic representation of the two methods with which to assemble a fission bomb. Source: Fastfission/ Public
domain
8.8.4 https://chem.libretexts.org/@go/page/372956
Figure 8.8.1 : Design of a nuclear power plant with a pressurized water reactor (PWR).
1. reactor block, 2. cooling tower, 3. reactor, 4. control rod, 5. support for pressure, 6. steam generator, 7. fuel element, 8. turbine, 9.
generator, 10. transformer, 11. condenser, 12. gaseous, 13. liquid, 14.air, 15.air (humid), 16. river, 17. cooling-water circulation, 18.
primary circuit, 19. secondary circuit, 20. water vapor, 21. pump. Source: Steffen Kuntoff / CC BY-SA 2.0 DE
(https://creativecommons.org/licenses...2.0/de/deed.en)
The reactor core is surrounded by a coolant and moderator, which may be regular water (H2O) or heavy water (D2O), or some other
material. It is called coolant because it takes away the heat produced by the fission reaction and moderator because it also slows
down the fast neutrons released in fission. The slow neutrons are more effective for the fission reaction. The coolant transfers its
heat to a steam generator. All of these components are housed in a containment building to keep the radioactivity contained, even if
a nuclear accident happens. The steam is used to drive a turbine for electricity production, and the condensed water is cooled using
water from the cooling tower and returned for reuse, as illustrated in Fig. 8.8.6.
Nuclear power plants produce about 20% of the electricity in the USA. Fig. 8.8.7 shows a nuclear power plant in Arkansas. In
some countries, the contribution of nuclear power to the total electricity is much higher, e.g., about 70% of electricity is produced
by nuclear power plants in France. Nuclear power is considered an energy source of the future, or at least for a transition from
fossil fuel to the next major energy source when the fossil fuel reserves will exhaust.
8.8.5 https://chem.libretexts.org/@go/page/372956
Figure 8.8.1 : Russellville nuclear power plant in Arkansas, showing a cooling tower in the front with non-radioactive steam
evaporating from it. Source: Edibobb, CC BY 3.0 <https://creativecommons.org/licenses/by/3.0>, via Wikimedia Commons
The fission products in the nuclear power plants become radioactive waste that needs to be stored for at least 10 half-lives to reach
an acceptable radioactivity level. Based on the 28.8 years half-life of strontium-90 which is the long-lived and dangerous product in
nuclear waste, a storage time of 300 years is needed. Handling radioactive waste is an issue of concern associated with nuclear
power production. Plutonium-239 is another radioactive isotope produced from neutron absorption by uranium-238. Uranium-238
is a major constituent (99.28%) in the natural uranium along with fissile uranium-235 (0.71%). Plutonium-239 has a long half-life
(24,000 Years) but it can be extracted in a fuel reprocessing plant and used as a nuclear fuel.
8.8: Nuclear fusion and fission is shared under a Public Domain license and was authored, remixed, and/or curated by Muhammad Arif Malik.
8.8.6 https://chem.libretexts.org/@go/page/372956
Index
B
buffer
6.8: pH Buffers
Glossary
Sample Word 1 | Sample Definition 1
Detailed Licensing
Overview
Title: Introduction to General Chemistry (Malik)
Webpages: 79
All licenses found:
Public Domain: 88.6% (70 pages)
Undeclared: 11.4% (9 pages)
By Page
Introduction to General Chemistry (Malik) — Public 3.5: Naming binary covalent compounds — Public
Domain Domain
Front Matter — Undeclared 3.6: Lewis structures of molecules — Public Domain
TitlePage — Undeclared 3.7: Molecular shapes –Valence shell electron pair
InfoPage — Undeclared repulsion (VSEPR) theory — Public Domain
Table of Contents — Undeclared 3.8: Polarity of molecules — Public Domain
Dedication — Public Domain 3.9: Intramolecular forces and intermolecular forces
Licensing — Undeclared — Public Domain
1: Matter energy and their measurements — Public 4: Stoichiometry –the quantification of chemical
Domain reactions — Public Domain
1.1: Matter and energy — Public Domain 4.1: Stoichiometry — Public Domain
1.2: What is chemistry? — Public Domain 4.2: The mole — Public Domain
1.3: Measurements — Public Domain 4.3: Chemical reaction — Public Domain
1.4: Significant Figures — Public Domain 4.4: Patterns of chemical reactions — Public Domain
1.5: Unit conversions — Public Domain 4.5: Oxidation-reduction reactions — Public Domain
1.6: Equations and graphs — Public Domain 4.6: Energetics of chemical reactions — Public
1.7: Density and specific gravity measurements — Domain
Public Domain 4.7: Stoichiometric calculations — Public Domain
1.8: Heat and its measurements — Public Domain 5: Solutions — Public Domain
1.9: Heat and changes in physical states of matter — 5.1: Introduction to solution — Public Domain
Public Domain 5.2: Solubility — Public Domain
2: Elements — Public Domain 5.3: Electrolytes — Public Domain
5.4: Concentration of solutions — Public Domain
2.1: Dalton’s atomic theory — Public Domain
5.5: Osmosis — Public Domain
2.2: Subatomic particles and a modern view of an
atom — Public Domain 6: Acids and bases — Public Domain
2.3: Atoms of elements — Public Domain 6.1: What is an acid and a base? — Public Domain
2.4: The periodic table — Public Domain 6.2: Brønsted–Lowry acids and bases — Public
2.5: Electrons in atoms — Public Domain Domain
2.6: The periodic trends in properties of the elements 6.3: Strength of acids and bases — Public Domain
— Public Domain 6.4: Acid-base equilibrium — Public Domain
3: Compounds — Public Domain 6.5: Dissociation of water — Public Domain
3.1: Bonding in compounds — Public Domain 6.6: The pH — Public Domain
3.2: Naming binary ionic compounds — Public 6.7: Acid-base reactions — Public Domain
Domain 6.8: pH Buffers — Public Domain
3.3: Polyatomic ions and their compounds — Public 7: Gases — Public Domain
Domain 7.1: Characteristics of gases — Public Domain
3.4: Naming acids — Public Domain 7.2: The pressure-volume relationship — Public
Domain
1 https://chem.libretexts.org/@go/page/374367
7.3: The temperature-volume relationship — Public 8.2: Radioactivity — Public Domain
Domain 8.3: Half-life of radioisotopes — Public Domain
7.4: The pressure-temperature relationship — Public 8.4: Radiation measurements — Public Domain
Domain 8.5: Ionizing radiation exposures — Public Domain
7.5: The combined gas law — Public Domain 8.6: Medical uses of radioisotopes — Public Domain
7.6: The volume-amount relationship — Public 8.7: Making radioisotopes for medical uses — Public
Domain Domain
7.7: Ideal gas law — Public Domain 8.8: Nuclear fusion and fission — Public Domain
7.8: Dalton’s law of partial pressure — Public Back Matter — Undeclared
Domain Index — Undeclared
8: Nuclear chemistry — Public Domain Glossary — Undeclared
8.1: Introduction to nuclear chemistry — Public Detailed Licensing — Undeclared
Domain
2 https://chem.libretexts.org/@go/page/374367