Oregon Institute of Technology: Che 202 General Chemistry Ii
Oregon Institute of Technology: Che 202 General Chemistry Ii
TECHNOLOGY: CHE
202
GENERAL CHEMISTRY II
Addie Clark
Oregon Institute of Technology
Oregon Institute of Technology
Oregon Institute of Technology: CHE 202
General Chemistry II
Addie Clark
This text is disseminated via the Open Education Resource (OER) LibreTexts Project (https://LibreTexts.org) and like the hundreds
of other texts available within this powerful platform, it is freely available for reading, printing and "consuming." Most, but not all,
pages in the library have licenses that may allow individuals to make changes, save, and print this book. Carefully
consult the applicable license(s) before pursuing such effects.
Instructors can adopt existing LibreTexts texts or Remix them to quickly build course-specific resources to meet the needs of their
students. Unlike traditional textbooks, LibreTexts’ web based origins allow powerful integration of advanced features and new
technologies to support learning.
The LibreTexts mission is to unite students, faculty and scholars in a cooperative effort to develop an easy-to-use online platform
for the construction, customization, and dissemination of OER content to reduce the burdens of unreasonable textbook costs to our
students and society. The LibreTexts project is a multi-institutional collaborative venture to develop the next generation of open-
access texts to improve postsecondary education at all levels of higher learning by developing an Open Access Resource
environment. The project currently consists of 14 independently operating and interconnected libraries that are constantly being
optimized by students, faculty, and outside experts to supplant conventional paper-based books. These free textbook alternatives are
organized within a central environment that is both vertically (from advance to basic level) and horizontally (across different fields)
integrated.
The LibreTexts libraries are Powered by MindTouch® and are supported by the Department of Education Open Textbook Pilot
Project, the UC Davis Office of the Provost, the UC Davis Library, the California State University Affordable Learning Solutions
Program, and Merlot. This material is based upon work supported by the National Science Foundation under Grant No. 1246120,
1525057, and 1413739. Unless otherwise noted, LibreTexts content is licensed by CC BY-NC-SA 3.0.
Any opinions, findings, and conclusions or recommendations expressed in this material are those of the author(s) and do not
necessarily reflect the views of the National Science Foundation nor the US Department of Education.
Have questions or comments? For information about adoptions or adaptions contact [email protected]. More information on our
activities can be found via Facebook (https://facebook.com/Libretexts), Twitter (https://twitter.com/libretexts), or our blog
(http://Blog.Libretexts.org).
1 https://chem.libretexts.org/@go/page/169559
Unit 6: Molecular Polarity
6.1: Electronegativity and Polarity
6.1: Electronegativity and Polarity (Problems)
6.2: Molecular Shape and Polarity
6.2: Molecular Shape and Polarity (Problems)
6.3: Intermolecular Forces
6.3: Intermolecular Forces (Problems)
Unit 9: Semiconductors
9.1: Bond Types and Molecular Orbital Theory
9.1: Bond Types and Molecular Orbital Theory (Problems)
9.2: The Solid State of Matter
9.2: The Solid State of Matter (Problems)
9.3: Structure and General Properties of the Metalloids
Index
Glossary
Detailed Licensing
2 https://chem.libretexts.org/@go/page/169559
Licensing
A detailed breakdown of this resource's licensing can be found in Back Matter/Detailed Licensing.
1 https://chem.libretexts.org/@go/page/417506
CHAPTER OVERVIEW
Unit 1: The Quantum World
Unit Objectives
By the end of this unit, you should be able to:
Explain the basic behavior of waves, including travelling waves and standing waves
Describe the wave nature of light
Use appropriate equations to calculate related light-wave properties such as period, frequency, wavelength, and energy
Distinguish between line and continuous emission spectra
Describe the particle nature of light
Extend the concept of wave–particle duality that was observed in electromagnetic radiation to matter as well
Calculate deBroglie wavelengths
Thumbnail: 3D views of some the 2s atomic orbitals showing probability density and phase. Image used with permission (CC BY-
AS 4.0; Geek3).
Unit 1: The Quantum World is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
1
1.1: The Nature of Light
Skills to Develop
Explain the basic behavior of waves, including traveling waves and standing waves
Describe the wave nature of light
Use appropriate equations to calculate related light-wave properties such as period, frequency, wavelength, and energy
Video 1.1.1 : Is light a particle or a wave? This section will explore the answers to this question.
The nature of light has been a subject of inquiry since antiquity. In the seventeenth century, Isaac Newton performed experiments
with lenses and prisms and was able to demonstrate that white light consists of the individual colors of the rainbow combined
together. Newton explained his optics findings in terms of a "corpuscular" view of light, in which light was composed of streams of
extremely tiny particles travelling at high speeds according to Newton's laws of motion. Others in the seventeenth century, such as
Christiaan Huygens, had shown that optical phenomena such as reflection and refraction could be equally well explained in terms
of light as waves travelling at high speed through a medium called "luminiferous aether" that was thought to permeate all space.
Early in the nineteenth century, Thomas Young demonstrated that light passing through narrow, closely spaced slits produced
interference patterns that could not be explained in terms of Newtonian particles but could be easily explained in terms of waves.
Later in the nineteenth century, after James Clerk Maxwell developed his theory of electromagnetic radiation and showed that light
was the visible part of a vast spectrum of electromagnetic waves, the particle view of light became thoroughly discredited. By the
end of the nineteenth century, scientists viewed the physical universe as roughly comprising two separate domains: matter
composed of particles moving according to Newton's laws of motion, and electromagnetic radiation consisting of waves governed
by Maxwell's equations. Today, these domains are referred to as classical mechanics and classical electrodynamics (or classical
electromagnetism). Although there were a few physical phenomena that could not be explained within this framework, scientists at
that time were so confident of the overall soundness of this framework that they viewed these aberrations as puzzling paradoxes
that would ultimately be resolved somehow within this framework. As we shall see, these paradoxes led to a contemporary
framework that intimately connects particles and waves at a fundamental level called wave-particle duality, which has superseded
the classical view.
Visible light and other forms of electromagnetic radiation play important roles in chemistry, since they can be used to infer the
energies of electrons within atoms and molecules. Much of modern technology is based on electromagnetic radiation. For example,
radio waves from a mobile phone, X-rays used by dentists, the energy used to cook food in your microwave, the radiant heat from
red-hot objects, and the light from your television screen are forms of electromagnetic radiation that all exhibit wavelike behavior.
Waves
1.1.1 https://chem.libretexts.org/@go/page/111596
Light Is Waves: Crash Course Physics #…
#…
1.1.2 https://chem.libretexts.org/@go/page/111596
Figure 1.1.1 : One-dimensional sinusoidal waves show the relationship among wavelength, frequency, and speed. The wave with
the shortest wavelength has the highest frequency. Amplitude is one-half the height of the wave from peak to trough.
The product of a wave's wavelength (λ) and its frequency (ν), λν, is the speed of the wave. Thus, for electromagnetic radiation in a
vacuum:
8 −1
c = 2.998 × 10 ms = λν (1.1.1)
Wavelength and frequency are inversely proportional: As the wavelength increases, the frequency decreases. The inverse
proportionality is illustrated in Figure 1.1.2. This figure also shows the electromagnetic spectrum, the range of all types of
electromagnetic radiation. Each of the various colors of visible light has specific frequencies and wavelengths associated with
them, and you can see that visible light makes up only a small portion of the electromagnetic spectrum. Because the technologies
developed to work in various parts of the electromagnetic spectrum are different, for reasons of convenience and historical legacies,
different units are typically used for different parts of the spectrum. For example, radio waves are usually specified as frequencies
(typically in units of MHz), while the visible region is usually specified in wavelengths (typically in units of nm or angstroms).
Figure 1.1.2 : Portions of the electromagnetic spectrum are shown in order of decreasing frequency and increasing wavelength.
Examples of some applications for various wavelengths include positron emission tomography (PET) scans, X-ray imaging, remote
controls, wireless Internet, cellular telephones, and radios. (credit “Cosmic ray": modification of work by NASA; credit “PET
scan": modification of work by the National Institute of Health; credit “X-ray": modification of work by Dr. Jochen Lengerke;
credit “Dental curing": modification of work by the Department of the Navy; credit “Night vision": modification of work by the
1.1.3 https://chem.libretexts.org/@go/page/111596
Department of the Army; credit “Remote": modification of work by Emilian Robert Vicol; credit “Cell phone": modification of
work by Brett Jordan; credit “Microwave oven": modification of work by Billy Mabray; credit “Ultrasound": modification of work
by Jane Whitney; credit “AM radio": modification of work by Dave Clausen)
The TeraHertz Region
Terahertz radiation is a region of the electromagnetic spectrum with frequencies of 0.3 to 3 THz (or 1 mm to 0.1 mm), and was
previously defined as microwave radio waves or the far IR. Because of its seat into beteen microwaves and IR, scientists have
begin refering to the region as the "TeraHertz gap" Due to the ability of several atmospheric gases to absorb energy in this
region of the spectrum, it is unsuitable for radio communications, but a lot of research into uses of technology in this area of the
spectrum have emerged over the last decade. To read more about it visit here or here.
Since c is expressed in meters per second, we must also convert 589 nm to meters.
8 −1 9
2.998 × 10 m s 1 × 10 nm
14 −1
ν =( )( ) = 5.09 × 10 s (1.1.3)
589 nm 1 m
Exercise 1.1.1
One of the frequencies used to transmit and receive cellular telephone signals in the United States is 850 MHz. What is the
wavelength in meters of these radio waves?
Answer
0.353 m = 35.3 cm
Wireless Communication
Many valuable technologies operate in the radio (3 kHz-300 GHz) frequency region of the electromagnetic spectrum. At the low
frequency (low energy, long wavelength) end of this region are AM (amplitude modulation) radio signals (540-2830 kHz) that can
travel long distances. FM (frequency modulation) radio signals are used at higher frequencies (87.5-108.0 MHz). In AM radio, the
1.1.4 https://chem.libretexts.org/@go/page/111596
information is transmitted by varying the amplitude of the wave (Figure 1.1.5). In FM radio, by contrast, the amplitude is constant
and the instantaneous frequency varies.
Figure 1.1.3 : Radio and cell towers are typically used to transmit long-wavelength electromagnetic radiation. Increasingly, cell
towers are designed to blend in with the landscape, as with the Tucson, Arizona, cell tower (right) disguised as a palm tree. (credit
left: modification of work by Sir Mildred Pierce; credit middle: modification of work by M.O. Stevens)
Other technologies also operate in the radio-wave portion of the electromagnetic spectrum. For example, 4G cellular telephone
signals are approximately 880 MHz, while Global Positioning System (GPS) signals operate at 1.228 and 1.575 GHz, local area
wireless technology (Wi-Fi) networks operate at 2.4 to 5 GHz, and highway toll sensors operate at 5.8 GHz. The frequencies
associated with these applications are convenient because such waves tend not to be absorbed much by common building materials.
Figure 1.1.4 : This schematic depicts how amplitude modulation (AM) and frequency modulation (FM) can be used to transmit a
radio wave.
One particularly characteristic phenomenon of waves results when two or more waves come into contact: They interfere with each
other. Figure 1.1.5 shows the interference patterns that arise when light passes through narrow slits closely spaced about a
wavelength apart. The fringe patterns produced depend on the wavelength, with the fringes being more closely spaced for shorter
wavelength light passing through a given set of slits. When the light passes through the two slits, each slit effectively acts as a new
source, resulting in two closely spaced waves coming into contact at the detector (the camera in this case). The dark regions in
Figure 1.1.5 correspond to regions where the peaks for the wave from one slit happen to coincide with the troughs for the wave
from the other slit (destructive interference), while the brightest regions correspond to the regions where the peaks for the two
waves (or their two troughs) happen to coincide (constructive interference). Likewise, when two stones are tossed close together
into a pond, interference patterns are visible in the interactions between the waves produced by the stones. Such interference
patterns cannot be explained by particles moving according to the laws of classical mechanics.
1.1.5 https://chem.libretexts.org/@go/page/111596
Figure 1.1.5 : Interference fringe patterns are shown for light passing through two closely spaced, narrow slits. The spacing of the
fringes depends on the wavelength, with the fringes being more closely spaced for the shorter-wavelength blue light. (credit:
PASCO)
Dorothy Hodgkin
Because the wavelengths of X-rays (10-10,000 picometers [pm]) are comparable to the size of atoms, X-rays can be used to
determine the structure of molecules. When a beam of X-rays is passed through molecules packed together in a crystal, the X-
rays collide with the electrons and scatter. Constructive and destructive interference of these scattered X-rays creates a specific
diffraction pattern. Calculating backward from this pattern, the positions of each of the atoms in the molecule can be determined
very precisely. One of the pioneers who helped create this technology was Dorothy Crowfoot Hodgkin.
She was born in Cairo, Egypt, in 1910, where her British parents were studying archeology. Even as a young girl, she was
fascinated with minerals and crystals. When she was a student at Oxford University, she began researching how X-ray
crystallography could be used to determine the structure of biomolecules. She invented new techniques that allowed her and her
students to determine the structures of vitamin B12, penicillin, and many other important molecules. Diabetes, a disease that
affects 382 million people worldwide, involves the hormone insulin. Hodgkin began studying the structure of insulin in 1934,
but it required several decades of advances in the field before she finally reported the structure in 1969. Understanding the
structure has led to better understanding of the disease and treatment options.
Not all waves are travelling waves. Standing waves (also known as stationary waves) remain constrained within some region of
space. As we shall see, standing waves play an important role in our understanding of the electronic structure of atoms and
molecules. The simplest example of a standing wave is a one-dimensional wave associated with a vibrating string that is held fixed
at its two end points. Figure 1.1.6 shows the four lowest-energy standing waves (the fundamental wave and the lowest three
harmonics) for a vibrating string at a particular amplitude. Although the string's motion lies mostly within a plane, the wave itself is
considered to be one dimensional, since it lies along the length of the string. The motion of string segments in a direction
perpendicular to the string length generates the waves and so the amplitude of the waves is visible as the maximum displacement of
the curves seen in Figure 1.1.6. The key observation from the figure is that only those waves having an integer number, n, of half-
wavelengths between the end points can form. A system with fixed end points such as this restricts the number and type of the
possible waveforms. This is an example of quantization, in which only discrete values from a more general set of continuous values
of some property are observed. Another important observation is that the harmonic waves (those waves displaying more than one-
half wavelength) all have one or more points between the two end points that are not in motion. These special points are nodes. The
energies of the standing waves with a given amplitude in a vibrating string increase with the number of half-wavelengths n. Since
the number of nodes is n – 1, the energy can also be said to depend on the number of nodes, generally increasing as the number of
nodes increases.
Figure 1.1.6 : A vibrating string shows some one-dimensional standing waves. Since the two end points of the string are held
fixed, only waves having an integer number of half-wavelengths can form. The points on the string between the end points that are
1.1.6 https://chem.libretexts.org/@go/page/111596
not moving are called the nodes.
An example of two-dimensional standing waves is shown in Figure 1.1.7 which shows the vibrational patterns on a flat surface.
Although the vibrational amplitudes cannot be seen like they could in the vibrating string, the nodes have been made visible by
sprinkling the drum surface with a powder that collects on the areas of the surface that have minimal displacement. For one-
dimensional standing waves, the nodes were points on the line, but for two-dimensional standing waves, the nodes are lines on the
surface (for three-dimensional standing waves, the nodes are two-dimensional surfaces within the three-dimensional volume).
Because of the circular symmetry of the drum surface, its boundary conditions (the drum surface being tightly constrained to the
circumference of the drum) result in two types of nodes: radial nodes that sweep out all angles at constant radii and, thus, are seen
as circles about the center, and angular nodes that sweep out all radii at constant angles and, thus, are seen as lines passing through
the center. The upper left image in Figure 1.1.7 shows two radial nodes, while the image in the lower right shows the vibrational
pattern associated with three radial nodes and two angular nodes.
Figure 1.1.7 : Two-dimensional standing waves can be visualized on a vibrating surface. The surface has been sprinkled with a
powder that collects near the nodal lines. There are two types of nodes visible: radial nodes (circles) and angular nodes (radii).
For a more animated video, check this link out.
Radial Nodes & Imogen Heap
You can watch the formation of various radial nodes below as singer Imogen Heap projects her voice across a kettle drum.
Video 1.1.4 : Singer Imogen Heap projects her voice across a kettle drum.
1.1.7 https://chem.libretexts.org/@go/page/111596
Blackbody Radiation and the Ultraviolet Catastrophe
The last few decades of the nineteenth century witnessed intense research activity in commercializing newly discovered electric
lighting. This required obtaining a better understanding of the distributions of light emitted from various sources being considered.
Artificial lighting is usually designed to mimic natural sunlight within the limitations of the underlying technology. Such lighting
consists of a range of broadly distributed frequencies that form a continuous spectrum. Figure 1.1.8 shows the wavelength
distribution for sunlight. The most intense radiation is in the visible region, with the intensity dropping off rapidly for shorter
wavelength ultraviolet (UV) light, and more slowly for longer wavelength infrared (IR) light.
Figure 1.1.8 : The spectral distribution (light intensity vs. wavelength) of sunlight reaches the Earth's atmosphere as UV light,
visible light, and IR light. The unabsorbed sunlight at the top of the atmosphere has a distribution that approximately matches the
theoretical distribution of a blackbody at 5250 °C, represented by the blue curve. (credit: modification of work by American Society
for Testing and Materials (ASTM) Terrestrial Reference Spectra for Photovoltaic Performance Evaluation)
In Figure 1.1.8, the solar distribution is compared to a representative distribution, called a blackbody spectrum, that corresponds to
a temperature of 5250 °C. The blackbody spectrum matches the solar spectrum quite well. A blackbody is a convenient, ideal
emitter that approximates the behavior of many materials when heated. It is “ideal” in the same sense that an ideal gas is a
convenient, simple representation of real gases that works well, provided that the pressure is not too high nor the temperature too
low. A good approximation of a blackbody that can be used to observe blackbody radiation is a metal oven that can be heated to
very high temperatures. The oven has a small hole allowing for the light being emitted within the oven to be observed with a
spectrometer so that the wavelengths and their intensities can be measured. Figure 1.1.8 shows the resulting curves for some
representative temperatures. Each distribution depends only on a single parameter: the temperature. The maxima in the blackbody
curves, λmax, shift to shorter wavelengths as the temperature increases, reflecting the observation that metals being heated to high
temperatures begin to glow a darker red that becomes brighter as the temperature increases, eventually becoming white hot at very
high temperatures as the intensities of all of the visible wavelengths become appreciable. This common observation was at the heart
of the first paradox that showed the fundamental limitations of classical physics that we will examine.
Physicists derived mathematical expressions for the blackbody curves using well-accepted concepts from the theories of classical
mechanics and classical electromagnetism. The theoretical expressions as functions of temperature fit the observed experimental
blackbody curves well at longer wavelengths, but showed significant discrepancies at shorter wavelengths. Not only did the
theoretical curves not show a peak, they absurdly showed the intensity becoming infinitely large as the wavelength became smaller,
which would imply that everyday objects at room temperature should be emitting large amounts of UV light. This became known
as the “ultraviolet catastrophe” because no one could find any problems with the theoretical treatment that could lead to such
unrealistic short-wavelength behavior. Finally, around 1900, Max Planck derived a theoretical expression for blackbody radiation
that fit the experimental observations exactly (within experimental error). Planck developed his theoretical treatment by extending
the earlier work that had been based on the premise that the atoms composing the oven vibrated at increasing frequencies (or
1.1.8 https://chem.libretexts.org/@go/page/111596
decreasing wavelengths) as the temperature increased, with these vibrations being the source of the emitted electromagnetic
radiation. But where the earlier treatments had allowed the vibrating atoms to have any energy values obtained from a continuous
set of energies (perfectly reasonable, according to classical physics), Planck found that by restricting the vibrational energies to
discrete values for each frequency, he could derive an expression for blackbody radiation that correctly had the intensity dropping
rapidly for the short wavelengths in the UV region.
E = nhν, n = 1, 2, 3, . . . (1.1.4)
The quantity h is a constant now known as Planck's constant, in his honor. Although Planck was pleased he had resolved the
blackbody radiation paradox, he was disturbed that to do so, he needed to assume the vibrating atoms required quantized energies,
which he was unable to explain. The value of Planck's constant is very small, 6.626 × 10−34 joule seconds (J s), which helps explain
why energy quantization had not been observed previously in macroscopic phenomena.
Figure 1.1.9 : Blackbody spectral distribution curves are shown for some representative temperatures.
1.1.9 https://chem.libretexts.org/@go/page/111596
Video 1.1.5 : An overview of the Ultraviolet Catastrophe.
Figure 1.1.10 : Photons with low frequencies do not have enough energy to cause electrons to be ejected via the photoelectric
effect. For any frequency of light above the threshold frequency, the kinetic energy of an ejected electron will increase linearly with
the energy of the incoming photon.
1.1.10 https://chem.libretexts.org/@go/page/111596
−34 8 −1
(6.626 × 10 J s )(2.998 × 10 m s )
E =
1 m
(640 nm ) ( )
9
10 nm
−19
E = 3.10 × 10 J
Exercise 1.1.2
The microwaves in an oven are of a specific frequency that will heat the water molecules contained in food. (This is why most
plastics and glass do not become hot in a microwave oven-they do not contain water molecules.) This frequency is about 3 × 109
Hz. What is the energy of one photon in these microwaves?
Answer
2 × 10−24 J
Exercise 1.1.3
Calculate the threshold energy in kJ/mol of electrons in aluminum, given that the lowest frequency photon for which the
photoelectric effect is observed is 9.87 × 10 H z .
14
Answer
3.94 kJ/mol
1.1.11 https://chem.libretexts.org/@go/page/111596
Quantum Mechanics - Part 1: Crash Cou…
Cou…
Summary
1.1.12 https://chem.libretexts.org/@go/page/111596
Preview of Section 1.2
Key Equations
c = λν
hc
E = hν = , where h = 6.626 × 10−34 J s
λ
Glossary
amplitude
extent of the displacement caused by a wave (for sinusoidal waves, it is one-half the difference from the peak height to the
trough depth, and the intensity is proportional to the square of the amplitude)
blackbody
idealized perfect absorber of all incident electromagnetic radiation; such bodies emit electromagnetic radiation in characteristic
continuous spectra called blackbody radiation
continuous spectrum
electromagnetic radiation given off in an unbroken series of wavelengths (e.g., white light from the sun)
electromagnetic radiation
energy transmitted by waves that have an electric-field component and a magnetic-field component
electromagnetic spectrum
range of energies that electromagnetic radiation can comprise, including radio, microwaves, infrared, visible, ultraviolet, X-
rays, and gamma rays; since electromagnetic radiation energy is proportional to the frequency and inversely proportional to the
wavelength, the spectrum can also be specified by ranges of frequencies or wavelengths
frequency (ν)
number of wave cycles (peaks or troughs) that pass a specified point in space per unit time
hertz (Hz)
the unit of frequency, which is the number of cycles per second, s−1
intensity
property of wave-propagated energy related to the amplitude of the wave, such as brightness of light or loudness of sound
interference pattern
pattern typically consisting of alternating bright and dark fringes; it results from constructive and destructive interference of
waves
1.1.13 https://chem.libretexts.org/@go/page/111596
node
any point of a standing wave with zero amplitude
photon
smallest possible packet of electromagnetic radiation, a particle of light
quantization
occurring only in specific discrete values, not continuous
standing wave
(also, stationary wave) localized wave phenomenon characterized by discrete wavelengths determined by the boundary
conditions used to generate the waves; standing waves are inherently quantized
wave
oscillation that can transport energy from one point to another in space
wavelength (λ)
distance between two consecutive peaks or troughs in a wave
wave-particle duality
term used to describe the fact that elementary particles including matter exhibit properties of both particles (including localized
position, momentum) and waves (including nonlocalization, wavelength, frequency)
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Crash Course Physics: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
Crash Course Astronomy: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
TED-Ed’s commitment to creating lessons worth sharing is an extension of TED’s mission of spreading great ideas. Within
TED-Ed’s growing library of TED-Ed animations, you will find carefully curated educational videos, many of which represent
collaborations between talented educators and animators nominated through the TED-Ed website.
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
1.1: The Nature of Light is shared under a CC BY license and was authored, remixed, and/or curated by LibreTexts.
1.1.14 https://chem.libretexts.org/@go/page/111596
1.1: The Nature of Light (Problems)
PROBLEM 1.1.1
An FM radio station found at 103.1 on the FM dial broadcasts at a frequency of 1.031 × 108 s−1 (103.1 MHz). What is the
wavelength of these radio waves in meters?
Answer
2.908 m
PROBLEM 1.1.2
FM-95, an FM radio station, broadcasts at a frequency of 9.51 × 107 s−1 (95.1 MHz). What is the wavelength of these radio
waves in meters?
Answer
3.15 m
Click here to see a video of the solution
Problem 1.1.2
PROBLEM 1.1.3
Light with a wavelength of 614.5 nm looks orange. What is the energy, in joules, per photon of this orange light? What is the
energy in eV (1 eV = 1.602 × 10−19 J)?
Answer
−19
3.233 × 10 J
2.018 eV
PROBLEM 1.1.4
Heated lithium atoms emit photons of light with an energy of 2.961 × 10−19 J. Calculate the frequency and wavelength of one of
these photons. What is the total energy in 1 mole of these photons? What is the color of the emitted light?
Answer
Frequency: 4.469 × 10 14
Hz
Wavelength: 6.709 × 10 −7
m = Red Light
Total energy: 1.783 × 10 5
J
1.1.1 https://chem.libretexts.org/@go/page/111604
Problem 1.1.4
PROBLEM 1.1.5
A photon of light produced by a surgical laser has an energy of 3.027 × 10−19 J. Calculate the frequency and wavelength of the
photon. What is the total energy in 1 mole of photons? What is the color of the emitted light?
Answer
ν = 4.568 × 1014 s; λ = 656.3 nm; Energy mol−1 = 1.823 × 105 J mol−1; red
PROBLEM 1.1.6
One of the radiographic devices used in a dentist's office emits an X-ray of wavelength 2.090 × 10−11 m. What is the energy, in
joules, and frequency of this X-ray?
Answer
E = 9.502 × 10−15 J; ν = 1.434 × 1019 s−1
Click here to see a video of the solution
Problem 1.1.6
PROBLEM 1.1.7
1.1.2 https://chem.libretexts.org/@go/page/111604
The eyes of certain reptiles pass a single visual signal to the brain when the visual receptors are struck by photons of a
wavelength of 850 nm. If a total energy of 3.15 × 10−14 J is required to trip the signal, what is the minimum number of photons
that must strike the receptor?
Answer
5
1.3 × 10 photons
PROBLEM 1.1.8
Answer the following questions about a Blu-ray laser:
a. The laser on a Blu-ray player has a wavelength of 405 nm. In what region of the electromagnetic spectrum is this radiation?
What is its frequency?
b. A Blu-ray laser has a power of 5 milliwatts (1 watt = 1 J s−1). How many photons of light are produced by the laser in 1
hour?
c. The ideal resolution of a player using a laser (such as a Blu-ray player), which determines how close together data can be
stored on a compact disk, is determined using the following formula: Resolution = 0.60(λ/NA), where λ is the wavelength of
the laser (in nm) and NA is the numerical aperture (dimensionless). Numerical aperture is a measure of the size of the spot of
light on the disk; the larger the NA, the smaller the spot. In a typical Blu-ray system, NA = 0.95. If the 405-nm laser is used
in a Blu-ray player, what is the closest that information can be stored on a Blu-ray disk?
d. The data density of a Blu-ray disk using a 405-nm laser is 1.5 × 107 bits mm−2. Disks have an outside diameter of 120 mm
and a hole of 15-mm diameter. How many data bits can be contained on the disk? If a Blu-ray disk can hold 9,400,000 pages
of text, how many data bits are needed for a typed page? (Hint: Determine the area of the disk that is available to hold data.
The area inside a circle is given by A = πr2, where the radius r is one-half of the diameter.)
Answer a
7.40 × 1014 s-1
Answer b
3.67 × 1019 photons
Answer c
255 nm
Answer d
1.67 × 1011 bits
17,765 bits/page
Problem 1.1.8
1.1.3 https://chem.libretexts.org/@go/page/111604
PROBLEM 1.1.9
What is the threshold frequency for sodium metal if a photon with frequency 6.66 × 1014 s−1 ejects a photon with 7.74 × 10−20 J
kinetic energy? Will the photoelectric effect be observed if sodium is exposed to orange light?
Answer
5.49 × 1014 s−1; no
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Feedback
Think one of the answers above is wrong? Let us know here.
1.1: The Nature of Light (Problems) is shared under a CC BY license and was authored, remixed, and/or curated by LibreTexts.
1.1.4 https://chem.libretexts.org/@go/page/111604
1.2: Atomic Spectroscopy and the deBroglie Wavelength
Skills to Develop
Distinguish between line and continuous emission spectra
Describe the particle nature of light
Extend the concept of wave–particle duality that was observed in electromagnetic radiation to matter as well
Calculate deBroglie wavelengths
Line Spectra
Video 1.2.1 : A brief review of how wavelength and frequency affect the colors of light.
Another paradox within the classical electromagnetic theory that scientists in the late nineteenth century struggled with concerned
the light emitted from atoms and molecules. When solids, liquids, or condensed gases are heated sufficiently, they radiate some of
the excess energy as light. Photons produced in this manner have a range of energies, and thereby produce a continuous spectrum in
which an unbroken series of wavelengths is present. Most of the light generated from stars (including our sun) is produced in this
fashion. You can see all the visible wavelengths of light present in sunlight by using a prism to separate them. As can be seen in
Figure 1.1.8 in the previous section, sunlight also contains UV light (shorter wavelengths) and IR light (longer wavelengths) that
can be detected using instruments but that are invisible to the human eye. Incandescent (glowing) solids such as tungsten filaments
in incandescent lights also give off light that contains all wavelengths of visible light. These continuous spectra can often be
approximated by blackbody radiation curves at some appropriate temperature, such as those shown in the previous section.
In contrast to continuous spectra, light can also occur as discrete or line spectra having very narrow line widths interspersed
throughout the spectral regions such as those shown in Figure 1.2.2. Exciting a gas at low partial pressure using an electrical
current, or heating it, will produce line spectra. Fluorescent light bulbs and neon signs operate in this way (Figure 1.2.1). Each
element displays its own characteristic set of lines, as do molecules, although their spectra are generally much more complicated.
1.2.1 https://chem.libretexts.org/@go/page/111606
Figure 1.2.1 : Neon signs operate by exciting a gas at low partial pressure using an electrical current. This sign shows the
elaborate artistic effects that can be achieved. (credit: Dave Shaver)
Each emission line consists of a single wavelength of light, which implies that the light emitted by a gas consists of a set of discrete
energies. For example, when an electric discharge passes through a tube containing hydrogen gas at low pressure, the H2 molecules
are broken apart into separate H atoms, and we see a blue-pink color. Passing the light through a prism produces a line spectrum,
indicating that this light is composed of photons of four visible wavelengths, as shown in Figure 1.2.2.
Figure 1.2.2 : Line spectra of select gas. Compare the two types of emission spectra: continuous spectrum of white light (top) and
the line spectra of the light from excited sodium, hydrogen, calcium, and mercury atoms.
The origin of discrete spectra in atoms and molecules was extremely puzzling to scientists in the late nineteenth century, since
according to classical electromagnetic theory, only continuous spectra should be observed. Even more puzzling, in 1885, Johann
Balmer was able to derive an empirical equation that related the four visible wavelengths of light emitted by hydrogen atoms to
whole integers. That equation is the following one, in which k is a constant:
1 1 1
=k( − ) , n = 3, 4, 5, 6 (1.2.1)
2
λ 4 n
Other discrete lines for the hydrogen atom were found in the UV and IR regions. Johannes Rydberg generalized Balmer's work and
developed an empirical formula that predicted all of hydrogen's emission lines, not just those restricted to the visible range, where,
n1 and n2 are integers, n1 < n2, and R is the Rydberg constant (1.097 × 107 m−1).
∞
1 1 1
= R∞ ( − ) (1.2.2)
2 2
λ n n
1 2
Even in the late nineteenth century, spectroscopy was a very precise science, and so the wavelengths of hydrogen were measured to
very high accuracy, which implied that the Rydberg constant could be determined very precisely as well. That such a simple
formula as the Rydberg formula could account for such precise measurements seemed astounding at the time, but it was the
eventual explanation for emission spectra by Neils Bohr in 1913 that ultimately convinced scientists to abandon classical physics
and spurred the development of modern quantum mechanics.
1.2.2 https://chem.libretexts.org/@go/page/111606
Light: Crash Course Astronomy #24
Figure 1.2.3 : Dr. Neil Degrasse Tyson. Photo Credit: NASA/Bill Ingalls
In his book, Astrophysics for People in a Hurry, Neil Degrasse Tyson talks about the origins of the universe and how we've used
atomic spectroscopy and line spectra to determine what elements are present in various celestial bodies. Read versions of these
essays by clicking the links below:
"Over the Rainbow" - A look at Spectroscopy and it's role in learning more about our universe.
"The Periodic Table of the Cosmos" - A look at the origin of the elements (from space!)
1.2.3 https://chem.libretexts.org/@go/page/111606
Figure 1.2.4 : An interference pattern on the water surface is formed by interacting waves. The waves are caused by reflection of
water from the rocks. (credit: modification of work by Sukanto Debnath)
As technological improvements allowed scientists to probe the microscopic world in greater detail, it became increasingly clear by
the 1920s that very small pieces of matter follow a different set of rules from those we observe for large objects. The
unquestionable separation of waves and particles was no longer the case for the microscopic world.
One of the first people to pay attention to the special behavior of the microscopic world was Louis de Broglie. He asked the
question: If electromagnetic radiation can have particle-like character, can electrons and other submicroscopic particles exhibit
wavelike character? In his 1925 doctoral dissertation, de Broglie extended the wave–particle duality of light that Einstein used to
resolve the photoelectric-effect paradox to material particles. He predicted that a particle with mass m and velocity v (that is, with
linear momentum p) should also exhibit the behavior of a wave with a wavelength value λ, given by this expression in which h is
the familiar Planck’s constant
h h
λ = = (1.2.3)
mv p
This is called the de Broglie wavelength. Unlike the other values of λ discussed in this chapter, the de Broglie wavelength is a
characteristic of particles and other bodies, not electromagnetic radiation (note that this equation involves velocity [v, m/s], not
frequency [ν, Hz]. Although these two symbols are identical, they mean very different things). Where Bohr had postulated the
electron as being a particle orbiting the nucleus in quantized orbits, de Broglie argued that Bohr’s assumption of quantization can
be explained if the electron is considered not as a particle, but rather as a circular standing wave such that only an integer number
of wavelengths could fit exactly within the orbit (Figure 1.2.5).
Figure 1.2.5 : If an electron is viewed as a wave circling around the nucleus, an integer number of wavelengths must fit into the
orbit for this standing wave behavior to be possible.
For a circular orbit of radius r, the circumference is 2πr, and so de Broglie’s condition is:
2πr = nλ (1.2.4)
with n = 1, 2, 3, . . ..
Shortly after de Broglie proposed the wave nature of matter, two scientists at Bell Laboratories, C. J. Davisson and L. H. Germer,
demonstrated experimentally that electrons can exhibit wavelike behavior by showing an interference pattern for electrons
travelling through a regular atomic pattern in a crystal. The regularly spaced atomic layers served as slits, as used in other
interference experiments. Since the spacing between the layers serving as slits needs to be similar in size to the wavelength of the
tested wave for an interference pattern to form, Davisson and Germer used a crystalline nickel target for their “slits,” since the
spacing of the atoms within the lattice was approximately the same as the de Broglie wavelengths of the electrons that they used.
Figure 1.2.6 shows an interference pattern.
1.2.4 https://chem.libretexts.org/@go/page/111606
Figure 1.2.6 : (a) The interference pattern for electrons passing through very closely spaced slits demonstrates that quantum
particles such as electrons can exhibit wavelike behavior. (b) The experimental results illustrated here demonstrate the wave–
particle duality in electrons. The electrons pass through very closely spaced slits, forming an interference pattern, with increasing
numbers of electrons being recorded from the left image to the right. With only a few electrons recorded, it is clear that the
electrons arrive as individual localized “particles,” but in a seemingly random pattern. As more electrons arrive, a wavelike
interference pattern begins to emerge. Note that the probability of the final electron location is still governed by the wave-type
distribution, even for a single electron, but it can be observed more easily if many electron collisions have been recorded.
The wave–particle duality of matter can be seen by observing what happens if electron collisions are recorded over a long period of
time. Initially, when only a few electrons have been recorded, they show clear particle-like behavior, having arrived in small
localized packets that appear to be random. As more and more electrons arrived and were recorded, a clear interference pattern that
is the hallmark of wavelike behavior emerged. Thus, it appears that while electrons are small localized particles, their motion does
not follow the equations of motion implied by classical mechanics, but instead it is governed by some type of a wave equation that
governs a probability distribution even for a single electron’s motion. Thus the wave–particle duality first observed with photons is
actually a fundamental behavior intrinsic to all quantum particles.
Video 1.2.3 : View the Dr. Quantum – Double Slit Experiment cartoon for an easy-to-understand description of wave–particle
duality and the associated experiments.
1.2.5 https://chem.libretexts.org/@go/page/111606
We can use de Broglie’s equation to solve this problem, but we first must do a unit conversion of Planck’s constant. You learned
earlier that 1 J = 1 kg m2/s2. Thus, we can write h = 6.626 × 10–34 J s as 6.626 × 10–34 kg m2/s.
h
λ =
mv
−34 2
6.626 × 10 kg m /s
=
−31 7
(9.109 × 10 kg)(1.000 × 10 m/s)
−11
= 7.274 × 10 m
This is a small value, but it is significantly larger than the size of an electron in the classical (particle) view. This size is the same
order of magnitude as the size of an atom. This means that electron wavelike behavior is going to be noticeable in an atom.
Exercise 1.2.1
Calculate the wavelength of a softball with a mass of 100 g traveling at a velocity of 35 m s–1, assuming that it can be modeled
as a single particle.
Answer
1.9 × 10–34 m.
We never think of a thrown softball having a wavelength, since this wavelength is so small it is impossible for our senses or any
known instrument to detect (strictly speaking, the wavelength of a real baseball would correspond to the wavelengths of its
constituent atoms and molecules, which, while much larger than this value, would still be microscopically tiny). The de Broglie
wavelength is only appreciable for matter that has a very small mass and/or a very high velocity.
Werner Heisenberg considered the limits of how accurately we can measure properties of an electron or other microscopic particles.
He determined that there is a fundamental limit to how accurately one can measure both a particle’s position and its momentum
simultaneously. The more accurately we measure the momentum of a particle, the less accurately we can determine its position at
that time, and vice versa. This is summed up in what we now call the Heisenberg uncertainty principle: It is fundamentally
impossible to determine simultaneously and exactly both the momentum and the position of a particle. For a particle of mass m
moving with velocity vx in the x direction (or equivalently with momentum px), the product of the uncertainty in the position, Δx,
ℏ h
and the uncertainty in the momentum, Δpx , must be greater than or equal to (recall that ℏ = , the value of Planck’s constant
2 2π
divided by 2π).
ℏ
Δx × Δpx = (Δx)(mΔv) ≥ (1.2.5)
2
This equation allows us to calculate the limit to how precisely we can know both the simultaneous position of an object and its
momentum. For example, if we improve our measurement of an electron’s position so that the uncertainty in the position (Δx) has a
value of, say, 1 pm (10–12 m, about 1% of the diameter of a hydrogen atom), then our determination of its momentum must have an
uncertainty with a value of at least
−34 2
h (1.055 × 10 kg m /s)
−23
[Δp = mΔv = ] = = 5 × 10 kg m/s. (1.2.6)
−12
(2Δx) (2 × 1 × 10 m)
The value of ħ is not large, so the uncertainty in the position or momentum of a macroscopic object like a baseball is too
insignificant to observe. However, the mass of a microscopic object such as an electron is small enough that the uncertainty can be
large and significant.
It should be noted that Heisenberg’s uncertainty principle is not just limited to uncertainties in position and momentum, but it also
links other dynamical variables. For example, when an atom absorbs a photon and makes a transition from one energy state to
ℏ
another, the uncertainty in the energy and the uncertainty in the time required for the transition are similarly related, as ΔE Δt ≥ .
2
As will be discussed later, even the vector components of angular momentum cannot all be specified exactly simultaneously.
Heisenberg’s principle imposes ultimate limits on what is knowable in science. The uncertainty principle can be shown to be a
consequence of wave–particle duality, which lies at the heart of what distinguishes modern quantum theory from classical
1.2.6 https://chem.libretexts.org/@go/page/111606
mechanics. Recall that the equations of motion obtained from classical mechanics are trajectories where, at any given instant in
time, both the position and the momentum of a particle can be determined exactly. Heisenberg’s uncertainty principle implies that
such a view is untenable in the microscopic domain and that there are fundamental limitations governing the motion of quantum
particles. This does not mean that microscopic particles do not move in trajectories, it is just that measurements of trajectories are
limited in their precision. In the realm of quantum mechanics, measurements introduce changes into the system that is being
observed.
Summary
Macroscopic objects act as particles. Microscopic objects (such as electrons) have properties of both a particle and a wave. Their
exact trajectories cannot be determined. The quantum mechanical model of atoms describes the three-dimensional position of the
electron in a probabilistic manner according to a mathematical function called a wavefunction, often denoted as ψ. Atomic
wavefunctions are also called orbitals. The squared magnitude of the wavefunction describes the distribution of the probability of
finding the electron in a particular region in space. Therefore, atomic orbitals describe the areas in an atom where electrons are
most likely to be found.
Key Equation
h h
λ = =
mv p
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Crash Course Physics: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
Crash Course Astronomy: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
1.2: Atomic Spectroscopy and the deBroglie Wavelength is shared under a CC BY license and was authored, remixed, and/or curated by
LibreTexts.
1.2.7 https://chem.libretexts.org/@go/page/111606
1.2: Atomic Spectroscopy and the deBroglie Wavelength (Problems)
PROBLEM 1.2.1
Assume the mass of an electron is 9.11 x 10-31 kg.
a. Calculate the wavelength of an electron traveling with a speed of 2.65 x 106 m/s.
b. What would the wavelength of the helium atom be if it were traveling at the same speed as the electron in part a (1 amu =
1.6606 x 10-24 g)?
Answer a
2.74 x 10-10 m
Answer b
3.77 x 10-14 m
PROBLEM 1.2.2
What is the de Broglie wavelength of a 46 g baseball traveling at 94 mph?
Answer
3.43 x 10-34 m
Problem 1.2.2
PROBLEM 1.2.3
Which of the following equations describe particle-like behavior? Which describe wavelike behavior?
a. c = λν
2
mν
b. E =
2
2
n a0
c. r =
Z
d. E = hν
h
e. λ =
mν
Answer
e. describes particle-like behavior
a&d describe wavelike behavior
1.2.1 https://chem.libretexts.org/@go/page/111608
PROBLEM 1.2.4
RGB color television and computer displays use cathode ray tubes that produce colors by mixing red, green, and blue light. If
we look at the screen with a magnifying glass, we can see individual dots turn on and off as the colors change. Using a spectrum
of visible light, determine the approximate wavelength of each of these colors. What is the frequency and energy of a photon of
each of these colors
Answer
Red: 660 nm; 4.54 × 1014 Hz; 3.01 × 10−19 J. Green: 520 nm; 5.77 × 1014 Hz; 3.82 × 10−19 J. Blue: 440 nm; 6.81 × 1014 Hz;
4.51 × 10−19 J.
PROBLEM 1.2.5
A bright violet line occurs at 435.8 nm in the emission spectrum of mercury vapor. What amount of energy, in joules, must be
released by an electron in a mercury atom to produce a photon of this light?
Answer
−19
4.56 × 10 J
PROBLEM 1.2.6
When rubidium ions are heated to a high temperature, two lines are observed in its line spectrum at wavelengths (a) 7.9 × 10−7
m and (b) 4.2 × 10−7 m. What are the frequencies of the two lines? What color do we see when we heat a rubidium compound?
Answer
The frequency of (a) would be 3.79 × 1014 s-1.
The frequency of (b) would be 7.13 × 1014 s-1.
Because (a) would be in the near-IR, the compound would appear purple/blue based on the wavelength of (b).
Problem 1.2.6
PROBLEM 1.2.7
The emission spectrum of cesium contains two lines whose frequencies are (a) 3.45 × 1014 Hz and (b) 6.53 × 1014 Hz. What are
the wavelengths and energies per photon of the two lines? What color are the lines?
Answer a
λ = 8.69 × 10−7 m; E = 2.29 × 10−19 J; red
Answer b
1.2.2 https://chem.libretexts.org/@go/page/111608
λ = 4.59 × 10−7 m; E = 4.33 × 10−19 J; blue
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Feedback
Think one of the answers above is wrong? Let us know here.
1.2: Atomic Spectroscopy and the deBroglie Wavelength (Problems) is shared under a CC BY license and was authored, remixed, and/or curated
by LibreTexts.
1.2.3 https://chem.libretexts.org/@go/page/111608
CHAPTER OVERVIEW
Unit 2: Electrons in Atoms
Unit Objectives
Use Coulomb's Law to explain electrostatic potential and bonding
Understand the general idea of the quantum mechanical description of electrons in an atom, and that it uses the notion of
three-dimensional wave functions, or orbitals, that define the distribution of probability to find an electron in a particular part
of space
List and describe traits of the four quantum numbers that form the basis for completely specifying the state of an electron in
an atom
Describe the Bohr model of the hydrogen atom
Use the Rydberg equation to calculate energies of light emitted or absorbed by hydrogen atoms
Unit 2: Electrons in Atoms is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
1
2.1: Coulomb's Law and the Electrostatic Potential
Skills to Develop
Use Coulomb's Law to explain electrostatic potential and bonding
Through the work of scientists in the late 18th century, the main features of the electrostatic force—the existence of two types of charge, the observation
that like charges repel, unlike charges attract, and the decrease of force with distance—were eventually refined, and expressed as a mathematical formula.
The mathematical formula for the electrostatic force is called Coulomb’s law after the French physicist Charles Coulomb (1736–1806), who performed
experiments and first proposed a formula to calculate it.
Figure 2.1.1 : This NASA image of Arp 87 shows the result of a strong gravitational attraction between two galaxies. In contrast, at the subatomic level,
the electrostatic attraction between two objects, such as an electron and a proton, is far greater than their mutual attraction due to gravity. (credit:
NASA/HST)
Following the work of Ernest Rutherford and his colleagues in the early twentieth century, the picture of atoms consisting of tiny dense nuclei surrounded
by lighter and even tinier electrons continually moving about the nucleus was well established. This picture was called the planetary model, since it pictured
the atom as a miniature “solar system” with the electrons orbiting the nucleus like planets orbiting the sun. The simplest atom is hydrogen, consisting of a
single proton as the nucleus about which a single electron moves. The electrostatic force attracting the electron to the proton depends only on the distance
between the two particles, based on Coulomb's Law:
m1 m2
Fgravity = G (2.1.1)
2
r
with
G is a gravitational constant
m1 and m are the masses of particle 1 and 2, respectively
2
The electrostatic force has the same form as the gravitational force between two mass particles except that the electrostatic force depends on the magnitudes
of the charges on the particles (+1 for the proton and −1 for the electron) instead of the magnitudes of the particle masses that govern the gravitational force.
m1 m2
Felectrostatic = k (2.1.2)
2
r
with
k is a constant
m1 and m are the masses of particle 1 and 2, respectively
2
The electrostatic force is a vector quantity and is expressed in units of newtons. The force is understood to be along the line joining the two charges. (Figure
2.1.2)
Although the formula for Coulomb’s law is simple, it was no mean task to prove it. The experiments Coulomb did, with the primitive equipment then
available, were difficult. Modern experiments have verified Coulomb’s law to great precision. For example, it has been shown that the force is inversely
proportional to distance between two objects squared (F ∝ 1/r ) to an accuracy of 1 part in 10 . No exceptions have ever been found, even at the small
2 16
Figure 2.1.2 : The magnitude of the electrostatic forceF between point charges q and q separated by a distance r is given by Coulomb’s law. Note that
1 2
Newton’s third law (every force exerted creates an equal and opposite force) applies as usual—the force on q is equal in magnitude and opposite in
1
direction to the force it exerts on q . (a) Like charges. (b) Unlike charges.
2
2.1.1 https://chem.libretexts.org/@go/page/119817
Since forces can be derived from potentials, it is convenient to work with potentials instead, since they are forms of energy. The electrostatic potential is
also called the Coulomb potential. Because the electrostatic potential has the same form as the gravitational potential, according to classical mechanics, the
equations of motion should be similar, with the electron moving around the nucleus in circular or elliptical orbits (hence the label “planetary” model of the
atom). Potentials of the form V(r) that depend only on the radial distance r are known as central potentials. Central potentials have spherical symmetry, and
so rather than specifying the position of the electron in the usual Cartesian coordinates (x, y, z), it is more convenient to use polar spherical coordinates
centered at the nucleus, consisting of a linear coordinate r and two angular coordinates, usually specified by the Greek letters theta (θ) and phi (Φ). These
coordinates are similar to the ones used in GPS devices and most smart phones that track positions on our (nearly) spherical earth, with the two angular
coordinates specified by the latitude and longitude, and the linear coordinate specified by sea-level elevation. Because of the spherical symmetry of central
potentials, the energy and angular momentum of the classical hydrogen atom are constants, and the orbits are constrained to lie in a plane like the planets
orbiting the sun. This classical mechanics description of the atom is incomplete, however, since an electron moving in an elliptical orbit would be
accelerating (by changing direction) and, according to classical electromagnetism, it should continuously emit electromagnetic radiation. This loss in orbital
energy should result in the electron’s orbit getting continually smaller until it spirals into the nucleus, implying that atoms are inherently unstable.
visualize coulomb's law with this simulation!
Coulomb's Law
Atomic Scale
Macro Scale
Summary
2.1.2 https://chem.libretexts.org/@go/page/119817
Electric Charge: Crash Course Physics #25
Glossary
Coulomb’s law
the mathematical equation calculating the electrostatic force vector between two charged particles
Coulomb force
another term for the electrostatic force
electrostatic force
the amount and direction of attraction or repulsion between two charged bodies
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley (Stephen F. Austin State
University) with contributing authors. Textbook content produced by OpenStax College is licensed under a Creative Commons Attribution License 4.0
license. Download for free at http://cnx.org/contents/[email protected]).
Paul Peter Urone (Professor Emeritus at California State University, Sacramento) and Roger Hinrichs (State University of New York, College at
Oswego) with Contributing Authors: Kim Dirks (University of Auckland) and Manjula Sharma (University of Sydney). This work is licensed by
OpenStax University Physics under a Creative Commons Attribution License (by 4.0).
Adelaide Clark, Oregon Institute of Technology
Crash Course Physics: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
2.1: Coulomb's Law and the Electrostatic Potential is shared under a CC BY license and was authored, remixed, and/or curated by LibreTexts.
2.1.3 https://chem.libretexts.org/@go/page/119817
2.1: Coulomb's Law and the Electrostatic Potential (Problems)
PROBLEM 2.1.1
Would you expect two like charges close together to have a relative potential energy above or below 0?
Answer
Like charges should have energies above 0.
PROBLEM 2.1.2
Would you expect two like charges far apart to have a relative potential energy above or below 0?
Answer
Like charges should have energies above 0.
PROBLEM 2.1.3
Would you expect two opposite charges far apart to have a relative potential energy above or below 0?
Answer
opposite charges should have energies below 0.
PROBLEM 2.1.4
Rank the following scenarios from HIGHEST (most positive) to LOWEST (most negative) relative potential energy.
a. +3, -2 charges, 100 angstroms apart
b. +1, -1 charges, 200 angstroms apart
c. +1, +1 charges, 200 angstroms apart
d. -3, -3 charges, 100 angstroms apart
Answer
d, c, b, a
Problem 2.1.4
2.1.1 https://chem.libretexts.org/@go/page/119818
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Feedback
Think one of the answers above is wrong? Let us know here.
2.1: Coulomb's Law and the Electrostatic Potential (Problems) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by LibreTexts.
2.1.2 https://chem.libretexts.org/@go/page/119818
2.2: Atomic Orbitals and Quantum Numbers
Skills to Develop
Understand the general idea of the quantum mechanical description of electrons in an atom, and that it uses the notion of three-dimensional wave
functions, or orbitals, that define the distribution of probability to find an electron in a particular part of space
List and describe traits of the four quantum numbers that form the basis for completely specifying the state of an electron in an atom
protons and the negative charges of the electrons. So the further away the electron is from the nucleus, the greater the energy it has.
ΔE = Efinal − Einitial
1 1
−18
= −2.18 × 10 ( − ) J
2 2
n n
f i
2.2.1 https://chem.libretexts.org/@go/page/122444
The values nf and ni are the final and initial energy states of the electron.
The principal quantum number is one of three quantum numbers used to characterize an orbital. An atomic orbital, which is distinct from an orbit, is a
general region in an atom within which an electron is most probable to reside. The quantum mechanical model specifies the probability of finding an
electron in the three-dimensional space around the nucleus and is based on solutions of the Schrödinger equation. In addition, the principal quantum
number defines the energy of an electron in a hydrogen or hydrogen-like atom or an ion (an atom or an ion with only one electron) and the general
region in which discrete energy levels of electrons in a multi-electron atoms and ions are located.
Another quantum number is l, the angular momentum quantum number. It is an integer that defines the shape of the orbital, and takes on the values, l =
0, 1, 2, …, n – 1. This means that an orbital with n = 1 can have only one value of l, l = 0, whereas n = 2 permits l = 0 and l = 1, and so on. The
principal quantum number defines the general size and energy of the orbital. The l value specifies the shape of the orbital. Orbitals with the same value
of l form a subshell. In addition, the greater the angular momentum quantum number, the greater is the angular momentum of an electron at this orbital.
Orbitals with l = 0 are called s orbitals (or the s subshells). The value l = 1 corresponds to the p orbitals. For a given n, p orbitals constitute a p subshell
(e.g., 3p if n = 3). The orbitals with l = 2 are called the d orbitals, followed by the f-, g-, and h-orbitals for l = 3, 4, 5, and there are higher values we will
not consider.
There are certain distances from the nucleus at which the probability density of finding an electron located at a particular orbital is zero. In other words,
the value of the wavefunction ψ is zero at this distance for this orbital. Such a value of radius r is called a radial node. The number of radial nodes in an
orbital is n – l – 1.
Figure 2.2.2 : The graphs show the probability (y axis) of finding an electron for the 1s, 2s, 3s orbitals as a function of distance from the nucleus.
2.2.2 https://chem.libretexts.org/@go/page/122444
Figure 2.2.3 : Shapes of s, p, d, and f orbitals.
If an electron has an angular momentum (l ≠ 0), then this vector can point in different directions. In addition, the z component of the angular
momentum can have more than one value. This means that if a magnetic field is applied in the z direction, orbitals with different values of the z
component of the angular momentum will have different energies resulting from interacting with the field. The magnetic quantum number, called ml,
specifies the z component of the angular momentum for a particular orbital. For example, for an s orbital, l = 0, and the only value of ml is zero. For p
orbitals, l = 1, and ml can be equal to –1, 0, or +1. Generally speaking, ml can be equal to –l, –(l – 1), …, –1, 0, +1, …, (l – 1), l. The total number of
possible orbitals with the same value of l (a subshell) is 2l + 1. Thus, there is one s-orbital for ml = 0, there are three p-orbitals for ml = 1, five d-
orbitals for ml = 2, seven f-orbitals for ml = 3, and so forth. The principal quantum number defines the general value of the electronic energy. The
angular momentum quantum number determines the shape of the orbital. And the magnetic quantum number specifies orientation of the orbital in
space, as can be seen in Figure 2.2.3.
Figure 2.2.4 : The chart shows the energies of electron orbitals in a multi-electron atom.
Figure 2.2.4 illustrates the energy levels for various orbitals. The number before the orbital name (such as 2s, 3p, and so forth) stands for the principal
quantum number, n. The letter in the orbital name defines the subshell with a specific angular momentum quantum number l = 0 for s orbitals, 1 for p
orbitals, 2 for d orbitals. Finally, there are more than one possible orbitals for l ≥ 1, each corresponding to a specific value of ml. In the case of a
hydrogen atom or a one-electron ion (such as He+, Li2+, and so on), energies of all the orbitals with the same n are the same. This is called a
degeneracy, and the energy levels for the same principal quantum number, n, are called degenerate energy levels. However, in atoms with more than
one electron, this degeneracy is eliminated by the electron–electron interactions, and orbitals that belong to different subshells have different energies.
Orbitals within the same subshell (for example ns, np, nd, nf, such as 2p, 3s) are still degenerate and have the same energy.
2.2.3 https://chem.libretexts.org/@go/page/122444
While the three quantum numbers discussed in the previous paragraphs work well for describing electron orbitals, some experiments showed that they
were not sufficient to explain all observed results. It was demonstrated in the 1920s that when hydrogen-line spectra are examined at extremely high
resolution, some lines are actually not single peaks but, rather, pairs of closely spaced lines. This is the so-called fine structure of the spectrum, and it
implies that there are additional small differences in energies of electrons even when they are located in the same orbital. These observations led
Samuel Goudsmit and George Uhlenbeck to propose that electrons have a fourth quantum number. They called this the spin quantum number, or ms.
The other three quantum numbers, n, l, and ml, are properties of specific atomic orbitals that also define in what part of the space an electron is most
likely to be located. Orbitals are a result of solving the Schrödinger equation for electrons in atoms. The electron spin is a different kind of property. It
is a completely quantum phenomenon with no analogues in the classical realm. In addition, it cannot be derived from solving the Schrödinger equation
and is not related to the normal spatial coordinates (such as the Cartesian x, y, and z). Electron spin describes an intrinsic electron “rotation” or
“spinning.” Each electron acts as a tiny magnet or a tiny rotating object with an angular momentum, even though this rotation cannot be observed in
terms of the spatial coordinates.
The magnitude of the overall electron spin can only have one value, and an electron can only “spin” in one of two quantized states. One is termed the α
1
state, with the z component of the spin being in the positive direction of the z axis. This corresponds to the spin quantum number m s = . The other is
2
1
called the β state, with the z component of the spin being negative and ms = − . Any electron, regardless of the atomic orbital it is located in, can
2
1 1
only have one of those two values of the spin quantum number. The energies of electrons having ms = − and ms = are different if an external
2 2
magnetic field is applied.
Figure 2.2.5 : Electrons with spin values \(±\ce{1/2}\) in an external magnetic field.
1
Figure 2.2.5 illustrates this phenomenon. An electron acts like a tiny magnet. Its moment is directed up (in the positive direction of the z axis) for the
2
spin quantum number and down (in the negative z direction) for the spin quantum number of − . A magnet has a lower energy if its magnetic moment
1
is aligned with the external magnetic field (the left electron) and a higher energy for the magnetic moment being opposite to the applied field. This is
1 1
why an electron with m s = has a slightly lower energy in an external field in the positive z direction, and an electron with m s =− has a slightly
2 2
higher energy in the same field. This is true even for an electron occupying the same orbital in an atom. A spectral line corresponding to a transition for
electrons from the same orbital but with different spin quantum numbers has two possible values of energy; thus, the line in the spectrum will show a
fine structure splitting.
2.2.4 https://chem.libretexts.org/@go/page/122444
The Pauli Exclusion Principle
An electron in an atom is completely described by four quantum numbers: n, l, ml, and ms. The first three quantum numbers define the orbital and the
fourth quantum number describes the intrinsic electron property called spin. An Austrian physicist Wolfgang Pauli formulated a general principle that
gives the last piece of information that we need to understand the general behavior of electrons in atoms. The Pauli exclusion principle can be
formulated as follows: No two electrons in the same atom can have exactly the same set of all the four quantum numbers. What this means is that
electrons can share the same orbital (the same set of the quantum numbers n, l, and ml), but only if their spin quantum numbers ms have different
1
values. Since the spin quantum number can only have two values (± ) , no more than two electrons can occupy the same orbital (and if two electrons
2
are located in the same orbital, they must have opposite spins). Therefore, any atomic orbital can be populated by only zero, one, or two electrons. The
properties and meaning of the quantum numbers of electrons in atoms are briefly summarized in Table 2.2.1.
Table 2.2.1: Quantum Numbers, Their Properties, and Significance
Name Symbol Allowed values Physical meaning
Exercise 2.2.1
Identify the subshell in which electrons with the following quantum numbers are found:
a. n = 3, l = 1;
b. n = 5, l = 3;
c. n = 2, l = 0.
Answer a
3p
Answer b
5f
Answer c
2s
2.2.5 https://chem.libretexts.org/@go/page/122444
1 orbital labeled 5s
3 orbitals labeled 5p
5 orbitals labeled 5d
7 orbitals labeled 5f
+9 orbitals labeled 5g
–––––––––––––––––––––
25 orbitals total
Again, each orbital holds two electrons, so 50 electrons can fit in this shell.
(c) The number of orbitals in any shell n will equal n2. There can be up to two electrons in each orbital, so the maximum number of electrons will be
2 × n2
Exercise 2.2.2
If a shell contains a maximum of 32 electrons, what is the principal quantum number, n?
Answer
n=4
4f
4 1
7 7 3
5d
Solution
The table can be completed using the following rules:
The orbital designation is nl, where l = 0, 1, 2, 3, 4, 5, … is mapped to the letter sequence s, p, d, f, g, h, …,
The ml degeneracy is the number of orbitals within an l subshell, and so is 2l + 1 (there is one s orbital, three p orbitals, five d orbitals, seven f
orbitals, and so forth).
The number of radial nodes is equal to n – l – 1.
4f 4 3 7 0
4p 4 1 3 2
7f 7 3 7 3
5d 5 2 5 2
Exercise 2.2.3
How many orbitals have l = 2 and n = 3?
Answer
The five degenerate 3d orbitals
Summary
2.2.6 https://chem.libretexts.org/@go/page/122444
Orbitals: Crash Course Chemistry #25
Glossary
angular momentum quantum number (l)
quantum number distinguishing the different shapes of orbitals; it is also a measure of the orbital angular momentum
wavefunction (ψ)
mathematical description of an atomic orbital that describes the shape of the orbital; it can be used to calculate the probability of finding the electron
at any given location in the orbital, as well as dynamical variables such as the energy and the angular momentum
subshell
set of orbitals in an atom with the same values of n and l
spin quantum number (ms)
1 1
number specifying the electron spin direction, either + or −
2 2
shell
set of orbitals with the same principal quantum number, n
s orbital
spherical region of space with high electron density, describes orbitals with l = 0. An electron in this orbital is called an s electron
quantum mechanics
field of study that includes quantization of energy, wave-particle duality, and the Heisenberg uncertainty principle to describe matter
principal quantum number (n)
quantum number specifying the shell an electron occupies in an atom
Pauli exclusion principle
specifies that no two electrons in an atom can have the same value for all four quantum numbers
p orbital
dumbbell-shaped region of space with high electron density, describes orbitals with l = 1. An electron in this orbital is called a p electron
magnetic quantum number (ml)
quantum number signifying the orientation of an atomic orbital around the nucleus; orbitals having different values of ml but the same subshell
value of l have the same energy (are degenerate), but this degeneracy can be removed by application of an external magnetic field
Heisenberg uncertainty principle
rule stating that it is impossible to exactly determine both certain conjugate dynamical properties such as the momentum and the position of a
particle at the same time. The uncertainty principle is a consequence of quantum particles exhibiting wave–particle duality
f orbital
multilobed region of space with high electron density, describes orbitals with l = 3. An electron in this orbital is called an f electron
2.2.7 https://chem.libretexts.org/@go/page/122444
electron density
a measure of the probability of locating an electron in a particular region of space, it is equal to the squared absolute value of the wave function ψ
d orbital
region of space with high electron density that is either four lobed or contains a dumbbell and torus shape; describes orbitals with l = 2. An electron
in this orbital is called a d electron
atomic orbital
mathematical function that describes the behavior of an electron in an atom (also called the wavefunction), it can be used to find the probability of
locating an electron in a specific region around the nucleus, as well as other dynamical variables
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley (Stephen F. Austin State
University) with contributing authors. Textbook content produced by OpenStax College is licensed under a Creative Commons Attribution License
4.0 license. Download for free at http://cnx.org/contents/[email protected]).
Adelaide Clark, Oregon Institute of Technology
Crash Course Physics: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
TED-Ed’s commitment to creating lessons worth sharing is an extension of TED’s mission of spreading great ideas. Within TED-Ed’s growing
library of TED-Ed animations, you will find carefully curated educational videos, many of which represent collaborations between talented
educators and animators nominated through the TED-Ed website.
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
2.2: Atomic Orbitals and Quantum Numbers is shared under a CC BY license and was authored, remixed, and/or curated by LibreTexts.
2.2.8 https://chem.libretexts.org/@go/page/122444
2.2: Atomic Orbitals and Quantum Numbers (Problems)
PROBLEM 2.2.1
What are the allowed values for each of the four quantum numbers: n, l, ml, and ms?
Answer
n: non-zero integer
l: 0 to n-1
ml : -l to l
1 −1
ms : or
2 2
PROBLEM 2.2.2
Describe the properties of an electron associated with each of the following four quantum numbers: n, l, ml, and ms.
Answer
n determines the general range for the value of energy and the probable distances that the electron can be from the nucleus. l
determines the shape of the orbital. ml determines the orientation of the orbitals of the same l value with respect to one
another. ms determines the spin of an electron
PROBLEM 2.2.3
Identify the subshell in which electrons with the following quantum numbers are found:
a. n = 2, l = 1
b. n = 4, l = 2
c. n = 6, l = 0
Answer a
2p
Answer b
4d
Answer c
6s
PROBLEM 2.2.4
Identify the subshell in which electrons with the following quantum numbers are found:
a. n = 3, l = 2
b. n = 1, l = 0
c. n = 4, l = 3
Answer a
3d
Answer b
1s
Answer c
2.2.1 https://chem.libretexts.org/@go/page/122447
4f
PROBLEM 2.2.5
Consider the orbitals shown here in outline.
a. What is the maximum number of electrons contained in an orbital of type (x)? Of type (y)? Of type (z)?
b. How many orbitals of type (x) are found in a shell with n = 2? How many of type (y)? How many of type (z)?
c. Write a set of quantum numbers for an electron in an orbital of type (x) in a shell with n = 4. Of an orbital of type (y) in a
shell with n = 2. Of an orbital of type (z) in a shell with n = 3.
d. What is the smallest possible n value for an orbital of type (x)? Of type (y)? Of type (z)?
e. What are the possible l and ml values for an orbital of type (x)? Of type (y)? Of type (z)?
Answer a
x. 2
y. 2
z. 2
Answer b
x. 1
y. 3
z. 0
Answer c
1
x. 4 0 0
2
1
y. 2 1 0
2
1
z. 3 2 0
2
Answer d
x. 1
y. 2
z. 3
Answer e
x. l = 0, ml = 0
y. l = 1, ml = –1, 0, or +1
z. l = 2, ml = –2, –1, 0, +1, +2
PROBLEM 2.2.6
2.2.2 https://chem.libretexts.org/@go/page/122447
How many electrons could be held in the second shell of an atom if the spin quantum number ms could have three values instead
of just two? (Hint: Consider the Pauli exclusion principle.)
Answer
12
PROBLEM 2.2.7
Write a set of quantum numbers for each of the electrons with an n of 4 in a Se atom.
Answer
n l ml s
1
4 0 0 +
2
1
4 0 0 −
2
1
4 1 −1 +
2
1
4 1 0 +
2
1
4 1 +1 +
2
1
4 1 −1 −
2
PROBLEM 2.2.8
Answer the following questions:
a. Without using quantum numbers, describe the differences between the shells, subshells, and orbitals of an atom.
b. How do the quantum numbers of the shells, subshells, and orbitals of an atom differ?
Answer a
shell: set of orbitals in the same energy level
subshell: set of orbitals in the same energy level and same shape (s, p, d, or f)
orbital: can hold up to 2 electrons
Answer b
shell: set of orbitals with same n
subshell: set of orbitals in an atom with the same values of n and l
orbital: shape defined by l quantum number
PROBLEM 2.2.9
Sketch the boundary surface of a pz and a py orbital. Be sure to show and label the axes.
Answer
2.2.3 https://chem.libretexts.org/@go/page/122447
PROBLEM 2.2.10
Sketch the px and s orbitals. Be sure to show and label the coordinates.
Answer
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Feedback
Think one of the answers above is wrong? Let us know here.
2.2: Atomic Orbitals and Quantum Numbers (Problems) is shared under a CC BY license and was authored, remixed, and/or curated by
LibreTexts.
2.2.4 https://chem.libretexts.org/@go/page/122447
2.3: Atomic Orbitals and the Bohr Model
Skills to Develop
Describe the Bohr model of the hydrogen atom
Use the Rydberg equation to calculate energies of light emitted or absorbed by hydrogen atoms
Understand the general idea of the quantum mechanical description of electrons in an atom, and that it uses the notion of
three-dimensional wave functions, or orbitals, that define the distribution of probability to find an electron in a particular part
of space
In this equation, h is Planck’s constant and Ei and Ef are the initial and final orbital energies, respectively. The absolute value of the
energy difference is used, since frequencies and wavelengths are always positive. Instead of allowing for continuous values for the
angular momentum, energy, and orbit radius, Bohr assumed that only discrete values for these could occur (actually, quantizing any
one of these would imply that the other two are also quantized). Bohr’s expression for the quantized energies is:
k
En = − (2.3.2)
2
n
with n = 1, 2, 3, . . .
In this expression, k is a constant comprising fundamental constants such as the electron mass and charge and Planck’s constant.
Inserting the expression for the orbit energies into the equation for ΔE gives
1 1 hc
ΔE = k ( − ) = (2.3.3)
2 2
n n λ
1 2
or
2.3.1 https://chem.libretexts.org/@go/page/119819
1 k 1 1
= ( − ) (2.3.4)
2 2
λ hc n n
1 2
−18
1 1
ΔE = E1 − E2 = 2.179 × 10 ( − ) (2.3.5)
2 2
n n
1 2
This formula should look familiar - we talked about it in the previous section when we talked about the principle quantum number.
The lowest few energy levels are shown in Figure 2.3.1. One of the fundamental laws of physics is that matter is most stable with
the lowest possible energy. Thus, the electron in a hydrogen atom usually moves in the n = 1 orbit, the orbit in which it has the
lowest energy. When the electron is in this lowest energy orbit, the atom is said to be in its ground electronic state (or simply
ground state). If the atom receives energy from an outside source, it is possible for the electron to move to an orbit with a higher n
value and the atom is now in an excited electronic state (or simply an excited state) with a higher energy. When an electron
transitions from an excited state (higher energy orbit) to a less excited state, or ground state, the difference in energy is emitted as a
photon. Similarly, if a photon is absorbed by an atom, the energy of the photon moves an electron from a lower energy orbit up to a
more excited one.
Figure 2.3.1 : Quantum numbers and energy levels in a hydrogen atom. The more negative the calculated value, the lower the
energy.
We can relate the energy of electrons in atoms to what we learned previously about energy. The law of conservation of energy says
that we can neither create nor destroy energy. Thus, if a certain amount of external energy is required to excite an electron from one
energy level to another, that same amount of energy will be liberated when the electron returns to its initial state (Figure 2.3.2). In
effect, an atom can “store” energy by using it to promote an electron to a state with a higher energy and release it when the electron
returns to a lower state. The energy can be released as one quantum of energy, as the electron returns to its ground state (say, from
n = 5 to n = 1 ), or it can be released as two or more smaller quanta as the electron falls to an intermediate state, then to the
ground state (say, from n = 5 to n = 4 , emitting one quantum, then to n = 1 , emitting a second quantum).
2.3.2 https://chem.libretexts.org/@go/page/119819
Figure 2.3.2 : The horizontal lines show the relative energy of orbits in the Bohr model of the hydrogen atom, and the vertical
arrows depict the energy of photons absorbed (left) or emitted (right) as electrons move between these orbits.
Since Bohr’s model involved only a single electron, it could also be applied to the single electron ions He+, Li2+, Be3+, and so
forth, which differ from hydrogen only in their nuclear charges, and so one-electron atoms and ions are collectively referred to as
hydrogen-like or hydrogenic atoms. The energy expression for hydrogen-like atoms is a generalization of the hydrogen atom
energy, in which Z is the nuclear charge (+1 for hydrogen, +2 for He, +3 for Li, and so on) and k has a value of 2.179 × 10 J . –18
2
kZ
En = − (2.3.6)
2
n
The sizes of the circular orbits for hydrogen-like atoms are given in terms of their radii by the following expression, in which a is o
2
n
r = a0 (2.3.7)
Z
The equation also shows us that as the electron’s energy increases (as n increases), the electron is found at greater distances from
the nucleus. This is implied by the inverse dependence on r in the Coulomb potential, since, as the electron moves away from the
nucleus, the electrostatic attraction between it and the nucleus decreases, and it is held less tightly in the atom. Note that as n gets
larger and the orbits get larger, their energies get closer to zero, and so the limits n ⟶ ∞ and r ⟶ ∞ imply that E = 0
corresponds to the ionization limit where the electron is completely removed from the nucleus. Thus, for hydrogen in the ground
state n = 1 , the ionization energy would be:
ΔE = En⟶∞ − E1 = 0 + k = k (2.3.8)
With three extremely puzzling paradoxes now solved (blackbody radiation, the photoelectric effect, and the hydrogen atom), and all
involving Planck’s constant in a fundamental manner, it became clear to most physicists at that time that the classical theories that
worked so well in the macroscopic world were fundamentally flawed and could not be extended down into the microscopic domain
of atoms and molecules. Unfortunately, despite Bohr’s remarkable achievement in deriving a theoretical expression for the Rydberg
constant, he was unable to extend his theory to the next simplest atom, He, which only has two electrons. Bohr’s model was
severely flawed, since it was still based on the classical mechanics notion of precise orbits, a concept that was later found to be
untenable in the microscopic domain, when a proper model of quantum mechanics was developed to supersede classical
mechanics.
2.3.3 https://chem.libretexts.org/@go/page/119819
energy, in joules, of the electron?
Solution
The energy of the electron is given by Equation 2.3.6:
2
−kZ
E = (2.3.9)
n2
Exercise 2.3.1
The electron in Example 2.3.1 in the n = 3 state is promoted even further to an orbit with n = 6 . What is its new energy?
Answer
TBD
1 1
−18
ΔE = E1 − E2 = 2.179 × 10 ( − ) (2.3.11)
2 2
n n
1 2
1 1
−18
ΔE = 2.179 × 10 ( − ) J (2.3.12)
2 2
4 6
1 1
−18
ΔE = 2.179 × 10 ( − ) J (2.3.13)
16 36
−20
ΔE = 7.566 × 10 J (2.3.14)
This energy difference is positive, indicating a photon enters the system (is absorbed) to excite the electron from the n = 4 orbit
up to the n = 6 orbit. The wavelength of a photon with this energy is found by the expression E = hcλ . Rearrangement gives:
hc
λ = (2.3.15)
E
From the figure of electromagnetic radiation, we can see that this wavelength is found in the infrared portion of the
electromagnetic spectrum.
Exercise 2.3.2
What is the energy in joules and the wavelength in meters of the photon produced when an electron falls from the n = 5 to the
n = 3 level in a H e ion (Z = 2 for H e )?
+ +
Answer
6.198 × 10
–19
J and 3.205 × 10
−7
m
2.3.4 https://chem.libretexts.org/@go/page/119819
Bohr’s model of the hydrogen atom provides insight into the behavior of matter at the microscopic level, but it is does not account
for electron–electron interactions in atoms with more than one electron. It does introduce several important features of all models
used to describe the distribution of electrons in an atom. These features include the following:
The energies of electrons (energy levels) in an atom are quantized, described by quantum numbers: integer numbers having only
specific allowed value and used to characterize the arrangement of electrons in an atom.
An electron’s energy increases with increasing distance from the nucleus.
The discrete energies (lines) in the spectra of the elements result from quantized electronic energies.
Of these features, the most important is the postulate of quantized energy levels for an electron in an atom. As a consequence, the
model laid the foundation for the quantum mechanical model of the atom. Bohr won a Nobel Prize in Physics for his contributions
to our understanding of the structure of atoms and how that is related to line spectra emissions.
location in space. This means that wavefunctions can be used to determine the distribution of the electron’s density with respect to
the nucleus in an atom. In the most general form, the Schrödinger equation can be written as:
^
H ψ = Eψ (2.3.16)
^
H is the Hamiltonian operator, a set of mathematical operations representing the total energy of the quantum particle (such as an
electron in an atom), ψ is the wavefunction of this particle that can be used to find the special distribution of the probability of
finding the particle, and E is the actual value of the total energy of the particle.
2.3.5 https://chem.libretexts.org/@go/page/119819
Schrödinger’s work, as well as that of Heisenberg and many other scientists following in their footsteps, is generally referred to as
quantum mechanics.
Video 2.3.3 : You may also have heard of Schrödinger because of his famous thought experiment. This story explains the concepts
of superposition and entanglement as related to a cat in a box with poison.
Video 2.3.4 : What can we learn about Quantum Mechanics from the cat experiment?
2.3.6 https://chem.libretexts.org/@go/page/119819
Key Equations
−18
1 1
ΔE = E1 − E2 = 2.179 × 10 ( − ) (2.3.17)
2 2
n n
1 2
Summary
Bohr incorporated Planck’s and Einstein’s quantization ideas into a model of the hydrogen atom that resolved the paradox of atom
stability and discrete spectra. The Bohr model of the hydrogen atom explains the connection between the quantization of photons
and the quantized emission from atoms. Bohr described the hydrogen atom in terms of an electron moving in a circular orbit about
a nucleus. He postulated that the electron was restricted to certain orbits characterized by discrete energies. Transitions between
these allowed orbits result in the absorption or emission of photons. When an electron moves from a higher-energy orbit to a more
stable one, energy is emitted in the form of a photon. To move an electron from a stable orbit to a more excited one, a photon of
energy must be absorbed. Using the Bohr model, we can calculate the energy of an electron and the radius of its orbit in any one-
electron system.
Glossary
excited state
state having an energy greater than the ground-state energy
ground state
state in which the electrons in an atom, ion, or molecule have the lowest energy possible
line spectrum
electromagnetic radiation emitted at discrete wavelengths by a specific atom (or atoms) in an excited state
quantum number
integer number having only specific allowed values and used to characterize the arrangement of electrons in an atom
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Crash Course Physics: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
TED-Ed’s commitment to creating lessons worth sharing is an extension of TED’s mission of spreading great ideas. Within
TED-Ed’s growing library of TED-Ed animations, you will find carefully curated educational videos, many of which represent
collaborations between talented educators and animators nominated through the TED-Ed website.
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
2.3: Atomic Orbitals and the Bohr Model is shared under a CC BY license and was authored, remixed, and/or curated by LibreTexts.
2.3.7 https://chem.libretexts.org/@go/page/119819
2.3: Atomic Orbitals and the Bohr Model (Problems)
PROBLEM 2.3.1
Why is the electron in a Bohr hydrogen atom bound less tightly when it's electron is in energy level 3 than when it is in energy
level 1?
Answer
An n of 3 indicated that the 1 electron in the hydrogen atom is in the third energy level, which is further from the nucleus
than the first energy level (n=1), and therefore will not be as tightly bound.
PROBLEM 2.3.2
The electron volt (eV) is a convenient unit of energy for expressing atomic-scale energies. It is the amount of energy that an
electron gains when subjected to a potential of 1 volt; 1 eV = 1.602 × 10 −19
J . Using the Bohr model, determine the energy,
in electron volts, of the photon produced when an electron in a hydrogen atom moves from the orbit with n = 5 to the orbit with
n = 2.
Answer
2.856 eV
PROBLEM 2.3.3
Using the Bohr model, determine the energy in joules of the photon produced when an electron in a He
+
ion moves from the
orbit with n = 5 to the orbit with n = 2.
Answer
-4.58 × 10−19 J
Problem 2.3.3
PROBLEM 2.3.4
Using the Bohr model, determine the energy in joules of the photon produced when an electron in a Li2+ ion moves from the
orbit with n = 2 to the orbit with n = 1.
Answer
1.471 × 10−17 J
PROBLEM 2.3.5
2.3.1 https://chem.libretexts.org/@go/page/119820
Consider a large number of hydrogen atoms with electrons randomly distributed in the n = 1, 2, 3, and 4 orbits
a. How many different wavelengths of light are emitted by these atoms as the electrons fall into lower-energy orbitals?
b. Calculate the lowest and highest energies of light produced by the transitions described in part (a).
c. Calculate the frequencies and wavelengths of the light produced by the transitions described in part (b).
Answer a
6 possible falls producing 6 wavelengths
Answer b
Highest: n=4 to n=1; -2.04 x 10-18 J
Lowest: n=4 to n=3; -1.06 x 10-19 J
Answer c
Highest: 9.73 x 10-8 m; 3.08 x 1015 s-1
Lowest: 1.87 x 10-6 m; 1.60 x 1014 s-1
Problem 2.3.5
PROBLEM 2.3.6
The spectra of hydrogen and of calcium are shown below. What causes the lines in these spectra? Why are the colors of the lines
different? Suggest a reason for the observation that the spectrum of calcium is more complicated than the spectrum of hydrogen.
Answer
Both involve a relatively heavy nucleus with electrons moving around it, although strictly speaking, the Bohr model works
only for one-electron atoms or ions. According to classical mechanics, the Rutherford model predicts a miniature “solar
system” with electrons moving about the nucleus in circular or elliptical orbits that are confined to planes. If the
2.3.2 https://chem.libretexts.org/@go/page/119820
requirements of classical electromagnetic theory that electrons in such orbits would emit electromagnetic radiation are
ignored, such atoms would be stable, having constant energy and angular momentum, but would not emit any visible light
(contrary to observation). If classical electromagnetic theory is applied, then the Rutherford atom would emit
electromagnetic radiation of continually increasing frequency (contrary to the observed discrete spectra), thereby losing
energy until the atom collapsed in an absurdly short time (contrary to the observed long-term stability of atoms). The Bohr
model retains the classical mechanics view of circular orbits confined to planes having constant energy and angular
momentum, but restricts these to quantized values dependent on a single quantum number, n. The orbiting electron in Bohr’s
model is assumed not to emit any electromagnetic radiation while moving about the nucleus in its stationary orbits, but the
atom can emit or absorb electromagnetic radiation when the electron changes from one orbit to another. Because of the
quantized orbits, such “quantum jumps” will produce discrete spectra, in agreement with observations.
PROBLEM 2.3.7
How are the Bohr model and the quantum mechanical model of the hydrogen atom similar? How are they different?
Answer
Both models have a central positively charged nucleus with electrons moving about the nucleus in accordance with the
Coulomb electrostatic potential. The Bohr model assumes that the electrons move in circular orbits that have quantized
energies, angular momentum, and radii that are specified by a single quantum number, n = 1, 2, 3, …, but this quantization is
an ad hoc assumption made by Bohr to incorporate quantization into an essentially classical mechanics description of the
atom. Bohr also assumed that electrons orbiting the nucleus normally do not emit or absorb electromagnetic radiation, but do
so when the electron switches to a different orbit. In the quantum mechanical model, the electrons do not move in precise
orbits (such orbits violate the Heisenberg uncertainty principle) and, instead, a probabilistic interpretation of the electron’s
position at any given instant is used, with a mathematical function ψ called a wavefunction that can be used to determine the
electron’s spatial probability distribution. These wavefunctions, or orbitals, are three-dimensional stationary waves that can
be specified by three quantum numbers that arise naturally from their underlying mathematics (no ad hoc assumptions
required): the principal quantum number, n (the same one used by Bohr), which specifies shells such that orbitals having the
same n all have the same energy and approximately the same spatial extent; the angular momentum quantum number l,
which is a measure of the orbital’s angular momentum and corresponds to the orbitals’ general shapes, as well as specifying
subshells such that orbitals having the same l (and n) all have the same energy; and the orientation quantum number m,
which is a measure of the z component of the angular momentum and corresponds to the orientations of the orbitals. The
Bohr model gives the same expression for the energy as the quantum mechanical expression and, hence, both properly
account for hydrogen’s discrete spectrum (an example of getting the right answers for the wrong reasons, something that
many chemistry students can sympathize with), but gives the wrong expression for the angular momentum (Bohr orbits
necessarily all have non-zero angular momentum, but some quantum orbitals [s orbitals] can have zero angular momentum).
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Feedback
Think one of the answers above is wrong? Let us know here.
2.3: Atomic Orbitals and the Bohr Model (Problems) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.
2.3.3 https://chem.libretexts.org/@go/page/119820
CHAPTER OVERVIEW
Unit 3: Periodic Patterns
Unit Objectives
Derive the predicted ground-state electron configurations of atoms
Identify and explain exceptions to predicted electron configurations for atoms and ions
Predict the charge of common metallic and nonmetallic elements, and write their electron configurations
Relate electron configurations to element classifications in the periodic table
Describe and explain the observed trends in atomic size of the elements
Describe and explain the observed trends in ionization energy of the elements
Describe and explain the observed trends in electron affinity and metallic character of the elements
To recognize the inverse relationship of ionization energies and electron affinities
Unit 3: Periodic Patterns is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
1
3.1: Electron Configurations
Skills to Develop
Derive the predicted ground-state electron configurations of atoms
Identify and explain exceptions to predicted electron configurations for atoms and ions
Predict the charge of common metallic and nonmetallic elements, and write their electron configurations
Relate electron configurations to element classifications in the periodic table
Having introduced the basics of atomic structure and quantum mechanics, we can use our understanding of quantum numbers to
determine how atomic orbitals relate to one another. This allows us to determine which orbitals are occupied by electrons in each
atom. The specific arrangement of electrons in orbitals of an atom determines many of the chemical properties of that atom.
Figure 3.1.1 : Generalized energy-level diagram for atomic orbitals in an atom with two or more electrons (not to scale).
Electrons in successive atoms on the periodic table tend to fill low-energy orbitals first. Thus, many students find it confusing that,
for example, the 5p orbitals fill immediately after the 4d, and immediately before the 6s. The filling order is based on observed
experimental results, and has been confirmed by theoretical calculations. As the principal quantum number, n, increases, the size of
the orbital increases and the electrons spend more time farther from the nucleus. Thus, the attraction to the nucleus is weaker and
the energy associated with the orbital is higher (less stabilized). But this is not the only effect we have to take into account. Within
each shell, as the value of l increases, the electrons are less penetrating (meaning there is less electron density found close to the
nucleus), in the order s > p > d > f. Electrons that are closer to the nucleus slightly repel electrons that are farther out, offsetting the
more dominant electron–nucleus attractions slightly (recall that all electrons have −1 charges, but nuclei have +Z charges). This
phenomenon is called shielding and will be discussed in more detail in the next section. Electrons in orbitals that experience more
shielding are less stabilized and thus higher in energy. For small orbitals (1s through 3p), the increase in energy due to n is more
significant than the increase due to l; however, for larger orbitals the two trends are comparable and cannot be simply predicted. We
will discuss methods for remembering the observed order.
The arrangement of electrons in the orbitals of an atom is called the electron configuration of the atom. We describe an electron
configuration with a symbol that contains three pieces of information ( Figure 3.1.2):
3.1.1 https://chem.libretexts.org/@go/page/119829
1. The number of the principal quantum shell, n,
2. The letter that designates the orbital type (the subshell, l), and
3. A superscript number that designates the number of electrons in that particular subshell.
For example, the notation 2p4 (read "two–p–four") indicates four electrons in a p subshell (l = 1) with a principal quantum number
(n) of 2. The notation 3d8 (read "three–d–eight") indicates eight electrons in the d subshell (i.e., l = 2) of the principal shell for
which n = 3.
Figure 3.1.2 : The diagram of an electron configuration specifies the subshell (n and l value, with letter symbol) and superscript
number of electrons.
Figure 3.1.3 : The arrow leads through each subshell in the appropriate filling order for electron configurations. This chart is
straightforward to construct. Simply make a column for all the s orbitals with each n shell on a separate row. Repeat for p, d, and f.
Be sure to only include orbitals allowed by the quantum numbers (no 1p or 2d, and so forth). Finally, draw diagonal lines from top
to bottom as shown.
Since the arrangement of the periodic table is based on the electron configurations, Figure 3.1.4 provides an alternative method for
determining the electron configuration. The filling order simply begins at hydrogen and includes each subshell as you proceed in
increasing Z order. For example, after filling the 3p block up to Ar, we see the orbital will be 4s (K, Ca), followed by the 3d
orbitals.
3.1.2 https://chem.libretexts.org/@go/page/119829
Figure 3.1.4 : This periodic table shows the electron configuration for each subshell. By “building up” from hydrogen, this table
can be used to determine the electron configuration for any atom on the periodic table.
We will now construct the ground-state electron configuration and orbital diagram for a selection of atoms in the first and second
periods of the periodic table. Orbital diagrams are pictorial representations of the electron configuration, showing the individual
orbitals and the pairing arrangement of electrons. We start with a single hydrogen atom (atomic number 1), which consists of one
proton and one electron. Referring to either Figure 3.1.3 or 3.1.4, we would expect to find the electron in the 1s orbital. By
1
convention, the m s =+ value is usually filled first. The electron configuration and the orbital diagram are:
2
Following hydrogen is the noble gas helium, which has an atomic number of 2. The helium atom contains two protons and two
1
electrons. The first electron has the same four quantum numbers as the hydrogen atom electron (n = 1, l = 0, ml = 0, ms = + ).
2
The second electron also goes into the 1s orbital and fills that orbital. The second electron has the same n, l, and ml quantum
1
numbers, but must have the opposite spin quantum number, ms = − . This is in accord with the Pauli exclusion principle: No
2
two electrons in the same atom can have the same set of four quantum numbers. For orbital diagrams, this means two arrows go in
3.1.3 https://chem.libretexts.org/@go/page/119829
each box (representing two electrons in each orbital) and the arrows must point in opposite directions (representing paired spins).
The electron configuration and orbital diagram of helium are:
An atom of the alkaline earth metal beryllium, with an atomic number of 4, contains four protons in the nucleus and four electrons
surrounding the nucleus. The fourth electron fills the remaining space in the 2s orbital.
An atom of boron (atomic number 5) contains five electrons. The n = 1 shell is filled with two electrons and three electrons will
occupy the n = 2 shell. Because any s subshell can contain only two electrons, the fifth electron must occupy the next energy level,
which will be a 2p orbital. There are three degenerate 2p orbitals (ml = −1, 0, +1) and the electron can occupy any one of these p
orbitals. When drawing orbital diagrams, we include empty boxes to depict any empty orbitals in the same subshell that we are
filling.
Carbon (atomic number 6) has six electrons. Four of them fill the 1s and 2s orbitals. The remaining two electrons occupy the 2p
subshell. We now have a choice of filling one of the 2p orbitals and pairing the electrons or of leaving the electrons unpaired in two
different, but degenerate, p orbitals. The orbitals are filled as described by Hund’s rule: the lowest-energy configuration for an atom
with electrons within a set of degenerate orbitals is that having the maximum number of unpaired electrons. Thus, the two electrons
in the carbon 2p orbitals have identical n, l, and ms quantum numbers and differ in their ml quantum number (in accord with the
Pauli exclusion principle). The electron configuration and orbital diagram for carbon are:
Nitrogen (atomic number 7) fills the 1s and 2s subshells and has one electron in each of the three 2p orbitals, in accordance with
Hund’s rule. These three electrons have unpaired spins. Oxygen (atomic number 8) has a pair of electrons in any one of the 2p
orbitals (the electrons have opposite spins) and a single electron in each of the other two. Fluorine (atomic number 9) has only one
2p orbital containing an unpaired electron. All of the electrons in the noble gas neon (atomic number 10) are paired, and all of the
orbitals in the n = 1 and the n = 2 shells are filled. The electron configurations and orbital diagrams of these four elements are:
3.1.4 https://chem.libretexts.org/@go/page/119829
The alkali metal sodium (atomic number 11) has one more electron than the neon atom. This electron must go into the lowest-
energy subshell available, the 3s orbital, giving a 1s22s22p63s1 configuration. The electrons occupying the outermost shell orbital(s)
(highest value of n) are called valence electrons, and those occupying the inner shell orbitals are called core electrons ( Figure
3.1.5). Since the core electron shells correspond to noble gas electron configurations, we can abbreviate electron configurations by
writing the noble gas that matches the core electron configuration, along with the valence electrons in a condensed format. For our
sodium example, the symbol [Ne] represents core electrons, (1s22s22p6) and our abbreviated or condensed configuration is [Ne]3s1.
Figure 3.1.5 : A core-abbreviated electron configuration (right) replaces the core electrons with the noble gas symbol whose
configuration matches the core electron configuration of the other element.
Similarly, the abbreviated configuration of lithium can be represented as [He]2s1, where [He] represents the configuration of the
helium atom, which is identical to that of the filled inner shell of lithium. Writing the configurations in this way emphasizes the
similarity of the configurations of lithium and sodium. Both atoms, which are in the alkali metal family, have only one electron in a
valence s subshell outside a filled set of inner shells.
1
Li : [He] 2s (3.1.1)
1
Na : [Ne] 3s
The alkaline earth metal magnesium (atomic number 12), with its 12 electrons in a [Ne]3s2 configuration, is analogous to its family
member beryllium, [He]2s2. Both atoms have a filled s subshell outside their filled inner shells. Aluminum (atomic number 13),
with 13 electrons and the electron configuration [Ne]3s23p1, is analogous to its family member boron, [He]2s22p1.
The electron configurations of silicon (14 electrons), phosphorus (15 electrons), sulfur (16 electrons), chlorine (17 electrons), and
argon (18 electrons) are analogous in the electron configurations of their outer shells to their corresponding family members
carbon, nitrogen, oxygen, fluorine, and neon, respectively, except that the principal quantum number of the outer shell of the
heavier elements has increased by one to n = 3. Figure 3.1.6 shows the lowest energy, or ground-state, electron configuration for
these elements as well as that for atoms of each of the known elements.
3.1.5 https://chem.libretexts.org/@go/page/119829
Figure 3.1.6 : This version of the periodic table shows the outer-shell electron configuration of each element. Note that down each
group, the configuration is often similar.
When we come to the next element in the periodic table, the alkali metal potassium (atomic number 19), we might expect that we
would begin to add electrons to the 3d subshell. However, all available chemical and physical evidence indicates that potassium is
like lithium and sodium, and that the next electron is not added to the 3d level but is, instead, added to the 4s level (Figure 3.1.3 or
3.1.4). As discussed previously, the 3d orbital with no radial nodes is higher in energy because it is less penetrating and more
shielded from the nucleus than the 4s, which has three radial nodes. Thus, potassium has an electron configuration of [Ar]4s1.
Hence, potassium corresponds to Li and Na in its valence shell configuration. The next electron is added to complete the 4s
subshell and calcium has an electron configuration of [Ar]4s2. This gives calcium an outer-shell electron configuration
corresponding to that of beryllium and magnesium.
Beginning with the transition metal scandium (atomic number 21), additional electrons are added successively to the 3d subshell.
This subshell is filled to its capacity with 10 electrons (remember that for l = 2 [d orbitals], there are 2l + 1 = 5 values of ml,
meaning that there are five d orbitals that have a combined capacity of 10 electrons). The 4p subshell fills next. Note that for three
series of elements, scandium (Sc) through copper (Cu), yttrium (Y) through silver (Ag), and lutetium (Lu) through gold (Au), a
total of 10 d electrons are successively added to the (n – 1) shell next to the n shell to bring that (n – 1) shell from 8 to 18 electrons.
For two series, lanthanum (La) through lutetium (Lu) and actinium (Ac) through lawrencium (Lr), 14 f electrons (l = 3, 2l + 1 = 7
ml values; thus, seven orbitals with a combined capacity of 14 electrons) are successively added to the (n – 2) shell to bring that
shell from 18 electrons to a total of 32 electrons.
3.1.6 https://chem.libretexts.org/@go/page/119829
electrons:
The last electron added is a 3p electron. Therefore, n = 3 and, for a p-type orbital, l = 1. The ml value could be –1, 0, or +1. The
1
three p orbitals are degenerate, so any of these ml values is correct. For unpaired electrons, convention assigns the value of +
2
1
for the spin quantum number; thus, m s =+ .
2
Exercise 3.1.1
Identify the atoms from the electron configurations given:
a. [Ar]4s23d5
b. [Kr]5s24d105p6
Answer a
Mn
Answer b
Xe
The periodic table can be a powerful tool in predicting the electron configuration of an element. However, we do find exceptions to
the order of filling of orbitals that are shown in Figure 3.1.3 or 3.1.4. For instance, the electron configurations of the transition
metals chromium (Cr; atomic number 24) and copper (Cu; atomic number 29), among others, are not those we would expect. In
general, such exceptions involve subshells with very similar energy, and small effects can lead to changes in the order of filling.
In the case of Cr and Cu, we find that half-filled and completely filled subshells apparently represent conditions of preferred
stability. This stability is such that an electron shifts from the 4s into the 3d orbital to gain the extra stability of a half-filled 3d
subshell (in Cr) or a filled 3d subshell (in Cu). Other exceptions also occur. For example, niobium (Nb, atomic number 41) is
predicted to have the electron configuration [Kr]5s24d3. Experimentally, we observe that its ground-state electron configuration is
actually [Kr]5s14d4. We can rationalize this observation by saying that the electron–electron repulsions experienced by pairing the
electrons in the 5s orbital are larger than the gap in energy between the 5s and 4d orbitals. There is no simple method to predict the
exceptions for atoms where the magnitude of the repulsions between electrons is greater than the small differences in energy
between subshells.
Video 3.1.1 : A trick for writing electron configurations based on the organization of the periodic table.
As described earlier, the periodic table arranges atoms based on increasing atomic number so that elements with the same chemical
properties recur periodically. When their electron configurations are added to the table (Figure 3.1.6), we also see a periodic
3.1.7 https://chem.libretexts.org/@go/page/119829
recurrence of similar electron configurations in the outer shells of these elements. Because they are in the outer shells of an atom,
valence electrons play the most important role in chemical reactions. The outer electrons have the highest energy of the electrons in
an atom and are more easily lost or shared than the core electrons. Valence electrons are also the determining factor in some
physical properties of the elements.
Elements in any one group (or column) have the same number of valence electrons; the alkali metals lithium and sodium each have
only one valence electron, the alkaline earth metals beryllium and magnesium each have two, and the halogens fluorine and
chlorine each have seven valence electrons. The similarity in chemical properties among elements of the same group occurs
because they have the same number of valence electrons. It is the loss, gain, or sharing of valence electrons that defines how
elements react.
It is important to remember that the periodic table was developed on the basis of the chemical behavior of the elements, well before
any idea of their atomic structure was available. Now we can understand why the periodic table has the arrangement it has—the
arrangement puts elements whose atoms have the same number of valence electrons in the same group. This arrangement is
emphasized in Figure 3.1.6, which shows in periodic-table form the electron configuration of the last subshell to be filled by the
Aufbau principle. The colored sections of Figure 3.1.6 show the three categories of elements classified by the orbitals being filled:
main group, transition, and inner transition elements. These classifications determine which orbitals are counted in the valence
shell, or highest energy level orbitals of an atom.
1. Main group elements (sometimes called representative elements) are those in which the last electron added enters an s or a p
orbital in the outermost shell, shown in blue and red in Figure 3.1.6. This category includes all the nonmetallic elements, as
well as many metals and the intermediate semimetallic elements. The valence electrons for main group elements are those with
the highest n level. For example, gallium (Ga, atomic number 31) has the electron configuration [Ar]4s23d104p1, which contains
three valence electrons (underlined). The completely filled d orbitals count as core, not valence, electrons.
2. Transition elements or transition metals. These are metallic elements in which the last electron added enters a d orbital. The
valence electrons (those added after the last noble gas configuration) in these elements include the ns and (n – 1) d electrons.
The official IUPAC definition of transition elements specifies those with partially filled d orbitals. Thus, the elements with
completely filled orbitals (Zn, Cd, Hg, as well as Cu, Ag, and Au in Figure 3.1.6) are not technically transition elements.
However, the term is frequently used to refer to the entire d block (colored yellow in Figure 3.1.6), and we will adopt this usage
in this textbook.
3. Inner transition elements are metallic elements in which the last electron added occupies an f orbital. They are shown in green
in Figure 3.1.6. The valence shells of the inner transition elements consist of the (n – 2)f, the (n – 1)d, and the ns subshells.
There are two inner transition series:
1. The lanthanide series: lanthanide (La) through lutetium (Lu)
2. The actinide series: actinide (Ac) through lawrencium (Lr)
Lanthanum and actinium, because of their similarities to the other members of the series, are included and used to name the series,
even though they are transition metals with no f electrons.
3.1.8 https://chem.libretexts.org/@go/page/119829
Solution
First, write out the electron configuration for each parent atom. We have chosen to show the full, unabbreviated configurations
to provide more practice for students who want it, but listing the core-abbreviated electron configurations is also acceptable.
Next, determine whether an electron is gained or lost. Remember electrons are negatively charged, so ions with a positive
charge have lost an electron. For main group elements, the last orbital gains or loses the electron. For transition metals, the last s
orbital loses an electron before the d orbitals.
a. Na: 1s22s22p63s1. Sodium cation loses one electron, so Na+: 1s22s22p63s1 = Na+: 1s22s22p6.
b. P: 1s22s22p63s23p3. Phosphorus trianion gains three electrons, so P3−: 1s22s22p63s23p6.
c. Al: 1s22s22p63s23p1. Aluminum dication loses two electrons Al2+: 1s22s22p63s23p1 = Al2+: 1s22s22p63s1.
d. Fe: 1s22s22p63s23p64s23d6. Iron(II) loses two electrons and, since it is a transition metal, they are removed from the 4s
orbital Fe2+: 1s22s22p63s23p64s23d6 = 1s22s22p63s23p63d6.
e. Sm: 1s22s22p63s23p64s23d104p65s24d105p66s24f6. Samarium trication loses three electrons. The first two will be lost from
the 6s orbital, and the final one is removed from the 4f orbital. Sm3+: 1s22s22p63s23p64s23d104p65s24d105p66s24f6 =
1s22s22p63s23p64s23d104p65s24d105p64f5.
Exercise 3.1.2
a. Which ion with a +2 charge has the electron configuration 1s22s22p63s23p63d104s24p64d5?
b. Which ion with a +3 charge has this configuration?
Answer a
Tc2+
Answer b
Ru3+
3.1.9 https://chem.libretexts.org/@go/page/119829
Example 3.1.3 : Determining the Electronic Structures of Cations
There are at least 14 elements categorized as “essential trace elements” for the human body. They are called “essential” because
they are required for healthy bodily functions, “trace” because they are required only in small amounts, and “elements” in spite
of the fact that they are really ions. Two of these essential trace elements, chromium and zinc, are required as Cr3+ and Zn2+.
Write the electron configurations of these cations.
Solution
First, write the electron configuration for the neutral atoms:
Zn: [Ar]3d104s2
Cr: [Ar]3d54s1
Next, remove electrons from the highest energy orbital. For the transition metals, electrons are removed from the s orbital first
and then from the d orbital. For the p-block elements, electrons are removed from the p orbitals and then from the s orbital. Zinc
is a member of group 12, so it should have a charge of 2+, and thus loses only the two electrons in its s orbital. Chromium is a
transition element and should lose its s electrons and then its d electrons when forming a cation. Thus, we find the following
electron configurations of the ions:
Zn2+: [Ar]3d10
Cr3+: [Ar]3d3
Exercise 3.1.3
Potassium and magnesium are required in our diet. Write the electron configurations of the ions expected from these elements.
Answer
K+: [Ar], Mg2+: [Ne]
Exercise 3.1.4
Write the electron configurations of a phosphorus atom and its negative ion. Give the charge on the anion.
Answer
P: [Ne]3s23p3
P3–: [Ne]3s23p6
In ionic compounds, electrons are transferred between atoms of different elements to form ions. But this is not the only way that
compounds can be formed. Atoms can also make chemical bonds by sharing electrons between each other. Such bonds are called
covalent bonds. Covalent bonds are formed between two atoms when both have similar tendencies to attract electrons to
3.1.10 https://chem.libretexts.org/@go/page/119829
themselves (i.e., when both atoms have identical or fairly similar ionization energies and electron affinities). For example, two
hydrogen atoms bond covalently to form an H2 molecule; each hydrogen atom in the H2 molecule has two electrons stabilizing it,
giving each atom the same number of valence electrons as the noble gas He.
Compounds that contain covalent bonds exhibit different physical properties than ionic compounds. Because the attraction between
molecules, which are electrically neutral, is weaker than that between electrically charged ions, covalent compounds generally have
much lower melting and boiling points than ionic compounds. In fact, many covalent compounds are liquids or gases at room
temperature, and, in their solid states, they are typically much softer than ionic solids. Furthermore, whereas ionic compounds are
good conductors of electricity when dissolved in water, most covalent compounds, being electrically neutral, are poor conductors
of electricity in any state.
Summary
Video 3.1.2 : An overview of the role of orbitals in electron configurations and how to write electron configurations.
The relative energy of the subshells determine the order in which atomic orbitals are filled (1s, 2s, 2p, 3s, 3p, 4s, 3d, 4p, and so on).
Electron configurations and orbital diagrams can be determined by applying the Pauli exclusion principle (no two electrons can
have the same set of four quantum numbers) and Hund’s rule (whenever possible, electrons retain unpaired spins in degenerate
orbitals).
Electrons in the outermost orbitals, called valence electrons, are responsible for most of the chemical behavior of elements. In the
periodic table, elements with analogous valence electron configurations usually occur within the same group. There are some
exceptions to the predicted filling order, particularly when half-filled or completely filled orbitals can be formed. The periodic table
can be divided into three categories based on the orbital in which the last electron to be added is placed: main group elements (s
and p orbitals), transition elements (d orbitals), and inner transition elements (f orbitals).
Glossary
Aufbau principle
procedure in which the electron configuration of the elements is determined by “building” them in order of atomic numbers,
adding one proton to the nucleus and one electron to the proper subshell at a time
core electron
electron in an atom that occupies the orbitals of the inner shells
electron configuration
electronic structure of an atom in its ground state given as a listing of the orbitals occupied by the electrons
Hund’s rule
every orbital in a subshell is singly occupied with one electron before any one orbital is doubly occupied, and all electrons in
singly occupied orbitals have the same spin
orbital diagram
3.1.11 https://chem.libretexts.org/@go/page/119829
pictorial representation of the electron configuration showing each orbital as a box and each electron as an arrow
valence electrons
electrons in the outermost or valence shell (highest value of n) of a ground-state atom; determine how an element reacts
valence shell
outermost shell of electrons in a ground-state atom; for main group elements, the orbitals with the highest n level (s and p
subshells) are in the valence shell, while for transition metals, the highest energy s and d subshells make up the valence shell
and for inner transition elements, the highest s, d, and f subshells are included
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
3.1: Electron Configurations is shared under a CC BY license and was authored, remixed, and/or curated by LibreTexts.
3.1.12 https://chem.libretexts.org/@go/page/119829
3.1: Electron Configurations (Problems)
PROBLEM 3.1.1
Using complete subshell notation (no abbreviations), predict the electron configuration of each of the following atoms:
a. C
b. P
c. V
d. Sb
e. Sm
Answer a
1s22s22p2
Answer b
1s22s22p63s23p3
Answer c
1s22s22p63s23p63d34s2
Answer d
1s22s22p63s23p63d104s24p64d105s25p3
Answer e
1s22s22p63s23p63d104s24p64d104f65s25p66s2.
PROBLEM 3.1.2
Using complete subshell notation, predict the electron configuration of each of the following atoms:
a. N
b. Si
c. Fe
d. Te
e. Tb
Answer a
1s22s22p3
Answer b
1s22s22p63s23p2
Answer c
1s22s22p63s23p64s23d6
Answer d
1s22s22p63s23p64s23d104p65s24d105p4
Answer e
1s22s22p63s23p64s23d104p65s24d105p66s24f9
PROBLEM 3.1.3
3.1.1 https://chem.libretexts.org/@go/page/119830
Use an orbital diagram to describe the electron configuration of the valence shell of each of the following atoms:
a. N
b. Si
c. Fe
d. Te
e. Mo
Answer a
Answer b
Answer c
Answer d
Answer e
PROBLEM 3.1.4
Using complete subshell notation (1s22s22p6, and so forth), predict the electron configurations of the following ions.
a. N3–
b. Ca2+
c. S–
d. Cs2+
e. Cr2+
f. Gd3+
Answer a
1s22s22p6
3.1.2 https://chem.libretexts.org/@go/page/119830
Answer b
1s22s22p63s23p6
Answer c
1s22s22p63s23p5
Answer d
1s22s22p63s23p64s23d104p65s24d105p5
Answer e
1s22s22p63s23p64s23d2
Answer f
1s22s22p63s23p64s23d104p65s24d105p66s24f5
Problem 3.1.4
PROBLEM 3.1.5
Which atom has the electron configuration: 1s22s22p63s23p64s23d104p65s24d2?
Answer
Zr
PROBLEM 3.1.6
Which atom has the electron configuration: 1s22s22p63s23p63d74s2?
Answer
Co
PROBLEM 3.1.7
a. Which ion with a +1 charge has the electron configuration 1s22s22p63s23p63d104s24p6?
b. Which ion with a –2 charge has this configuration?
Answer a
Rb+
Answer b
3.1.3 https://chem.libretexts.org/@go/page/119830
Se2−
PROBLEM 3.1.8
Which of the following atoms contains only three valence electrons: Li, B, N, F, Ne?
Answer
B
PROBLEM 3.1.9
Which of the following has two unpaired electrons?
a. Mg
b. Si
c. S
d. Both Mg and S
e. Both Si and S.
Answer
Although both (b) and (c) are correct, (e) encompasses both and is the best answer.
PROBLEM 3.1.10
Which atom would be expected to have a half-filled 6p subshell?
Answer
Bi
PROBLEM 3.1.11
Which atom would be expected to have a half-filled 4s subshell?
Answer
K
PROBLEM 3.1.12
In one area of Australia, the cattle did not thrive despite the presence of suitable forage. An investigation showed the cause to be
the absence of sufficient cobalt in the soil. Cobalt forms cations in two oxidation states, Co2+ and Co3+. Write the electron
structure of the two cations.
Answer
Co2+: 1s22s22p63s23p64s23d5
Co3+: 1s22s22p63s23p64s23d4
3.1.4 https://chem.libretexts.org/@go/page/119830
Problem 3.1.12
PROBLEM 3.1.13
Thallium was used as a poison in the Agatha Christie mystery story “The Pale Horse.” Thallium has two possible cationic
forms, +1 and +3. The +1 compounds are the more stable. Write the electron structure of the +1 cation of thallium.
Answer
1s22s22p63s23p63d104s24p64d105s25p66s24f145d10
PROBLEM 3.1.14
Write the electron configurations for the following atoms or ions:
a. B3+
b. O–
c. Cl3+
d. Ca2+
e. Ti
Answer a
1s2
Answer b
1s22s22p5
Answer c
1s22s22p63s23p2
Answer d
1s22s22p63s23p6
Answer e
1s22s22p63s23p64s23d2
PROBLEM 3.1.15
Cobalt–60 and iodine–131 are radioactive isotopes commonly used in nuclear medicine. How many protons, neutrons, and
electrons are in atoms of these isotopes? Write the complete electron configuration for each isotope.
Answer
Co has 27 protons, 27 electrons, and 33 neutrons: 1s22s22p63s23p64s23d7.
3.1.5 https://chem.libretexts.org/@go/page/119830
I has 53 protons, 53 electrons, and 78 neutrons: 1s22s22p63s23p63d104s24p64d105s25p5.
PROBLEM 3.1.16
Atoms of which group in the periodic table have a valence shell electron configuration of ns2np3?
Answer
15 (5A)
PROBLEM 3.1.17
Atoms of which group in the periodic table have a valence shell electron configuration of ns2?
Answer
2 (2A)
PROBLEM 3.1.18
Does a cation gain protons to form a positive charge or does it lose electrons?
Answer
The protons in the nucleus do not change during normal chemical reactions. Only the outer electrons move. Positive charges
form when electrons are lost.
PROBLEM 3.1.19
Iron(III) sulfate [Fe2(SO4)3] is composed of Fe3+ and SO 2−
4
ions. Explain why a sample of iron(III) sulfate is uncharged.
Answer
Two cations with a +3 charge give a total of +6 charge, while three anions of -2 charge give a total of -6 charge. +6-6=0, so
when these ions bond, the charges will cancel, leaving the resulting compound uncharged.
PROBLEM 3.1.20
Which of the following atoms would be expected to form negative ions in binary ionic compounds and which would be
expected to form positive ions: P, I, Mg, Cl, In, Cs, O, Pb, Co?
Answer
P, I, Cl, and O would form anions because they are nonmetals. Mg, In, Cs, Pb, and Co would form cations because they are
metals.
PROBLEM 3.1.21
Which of the following atoms would be expected to form negative ions in binary ionic compounds and which would be
expected to form positive ions: Br, Ca, Na, N, F, Al, Sn, S, Cd?
Answer
Anions: Br, N, F, S,
Cations: Ca, Na, Al, Sn (because it's a metal), Cd (because it is a metal)
PROBLEM 3.1.22
Predict the charge on the monatomic ions formed from the following atoms in binary ionic compounds:
a. P
3.1.6 https://chem.libretexts.org/@go/page/119830
b. Mg
c. Al
d. O
e. Cl
f. Cs
Answer a
P3–
Answer b
Mg2+
Answer c
Al3+
Answer d
O2–
Answer e
Cl–
Answer f
Cs+
PROBLEM 3.1.23
Predict the charge on the monatomic ions formed from the following atoms in binary ionic compounds:
a. I
b. Sr
c. K
d. N
e. S
f. In
Answer a
I-
Answer b
Sr2+
Answer c
K+
Answer d
N3-
Answer e
S2-
Answer f
In3+
3.1.7 https://chem.libretexts.org/@go/page/119830
PROBLEM 3.1.24
Write the noble gas electron configuration for each of the following ions:
a. As3–
b. I–
c. Be2+
d. Cd2+
e. O2–
f. Ga3+
g. Li+
h. N3–
i. Sn2+
j. Co2+
k. Fe2+
l. As3+
Answer a
[Ar]4s23d104p6
Answer b
[Kr]4d105s25p6
Answer c
1s2
Answer d
[Kr]4d10
Answer e
[He]2s22p6
Answer f
[Ar]3d10
Answer g
1s2
Answer h
[He]2s22p6
Answer i
[Kr]4d105s2
Answer j
[Ar]3d7
Answer k
[Ar]3d6
Answer l
[Ar]3d104s2
3.1.8 https://chem.libretexts.org/@go/page/119830
PROBLEM 3.1.25
Write out the full electron configuration for each of the following atoms and for the monatomic ion found in binary ionic
compounds containing the element:
a. Al
b. Br
c. Sr
d. Li
e. As
f. S
Answer a
Al: 1s22s22p63s23p1
Al3+: 1s22s22p6
Answer b
Br: 1s22s22p63s23p63d104s24p5
Br-: 1s22s22p63s23p63d104s24p6
Answer c
Sr: 1s22s22p63s23p63d104s24p65s2
Sr2+: 1s22s22p63s23p63d104s24p6
Answer d
Li: 1s22s1
Li+: 1s2
Answer e
As: 1s22s22p63s23p63d104s24p3
As3-: 1s22s22p63s23p63d104s24p6
Answer f
S: 1s22s22p63s23p4
S2-: 1s22s22p63s23p6
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Feedback
Think one of the answers above is wrong? Let us know here.
3.1: Electron Configurations (Problems) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
3.1.9 https://chem.libretexts.org/@go/page/119830
3.2: Trends in Size
Skills to Develop
Describe and explain the observed trends in atomic size of the elements
The elements in groups (vertical columns) of the periodic table exhibit similar chemical behavior. This similarity occurs because
the members of a group have the same number and distribution of electrons in their valence shells. However, there are also other
patterns in chemical properties on the periodic table. For example, as we move down a group, the metallic character of the atoms
increases. Oxygen, at the top of Group 16 (6A), is a colorless gas; in the middle of the group, selenium is a semiconducting solid;
and, toward the bottom, polonium is a silver-grey solid that conducts electricity.
As we go across a period from left to right, we add a proton to the nucleus and an electron to the valence shell with each successive
element. As we go down the elements in a group, the number of electrons in the valence shell remains constant, but the principal
quantum number increases by one each time. An understanding of the electronic structure of the elements allows us to examine
some of the properties that govern their chemical behavior. These properties vary periodically as the electronic structure of the
elements changes. They are (1) size (radius) of atoms and ions, (2) ionization energies, and (3) electron affinities.
Visualizing periodic trends
Explore visualizations of the periodic trends discussed in this section (and many more trends). With just a few clicks, you can
create three-dimensional versions of the periodic table showing atomic size or graphs of ionization energies from all measured
elements.
3.2.1 https://chem.libretexts.org/@go/page/119833
Figure 3.2.1 : (a) The radius of an atom is defined as one-half the distance between the nuclei in a molecule consisting of two
identical atoms joined by a covalent bond. The atomic radius for the halogens increases down the group as n increases. (b)
Covalent radii of the elements are shown to scale. The general trend is that radii increase down a group and decrease across a
period.
We know that as we scan down a group, the principal quantum number, n, increases by one for each element. Thus, the electrons
are being added to a region of space that is increasingly distant from the nucleus. Consequently, the size of the atom (and its
covalent radius) must increase as we increase the distance of the outermost electrons from the nucleus. This trend is illustrated for
the covalent radii of the halogens in Table 3.2.1 and Figure 3.2.1. The trends for the entire periodic table can be seen in Figure
3.2.2.
3.2.2 https://chem.libretexts.org/@go/page/119833
N
u
c
l
e
a
Atom Covalent radius (pm) r
c
h
a
r
g
e
+
F 64
9
+
Cl 99 1
7
+
Br 114 3
5
+
I 133 5
3
+
At 148 8
5
As shown in Figure 3.2.2, as we move across a period from left to right, we generally find that each element has a smaller covalent
radius than the element preceding it. This might seem counterintuitive because it implies that atoms with more electrons have a
smaller atomic radius. This can be explained with the concept of effective nuclear charge, Z . This is the pull exerted on a
ef f
specific electron by the nucleus, taking into account any electron–electron repulsions. For hydrogen, there is only one electron and
so the nuclear charge (Z) and the effective nuclear charge (Zeff) are equal. For all other atoms, the inner electrons partially shield the
outer electrons from the pull of the nucleus, and thus:
Zeff = Z − shielding (3.2.1)
Shielding is determined by the probability of another electron being between the electron of interest and the nucleus, as well as by
the electron–electron repulsions the electron of interest encounters. Core electrons are adept at shielding, while electrons in the
same valence shell do not block the nuclear attraction experienced by each other as efficiently. Thus, each time we move from one
element to the next across a period, Z increases by one, but the shielding increases only slightly. Thus, Zeff increases as we move
from left to right across a period. The stronger pull (higher effective nuclear charge) experienced by electrons on the right side of
the periodic table draws them closer to the nucleus, making the covalent radii smaller.
3.2.3 https://chem.libretexts.org/@go/page/119833
Figure 3.2.2 : Within each period, the trend in atomic radius decreases as Z increases; for example, from K to Kr. Within each
group (e.g., the alkali metals shown in purple), the trend is that atomic radius increases as Z increases.
Exercise 3.2.1
Give an example of an atom whose size is smaller than fluorine.
Answer
3.2.4 https://chem.libretexts.org/@go/page/119833
Ne or He
Figure 3.2.3 : The radius for a cation is smaller than the parent atom (Al), due to the lost electrons; the radius for an anion is
larger than the parent (S), due to the gained electrons.
Cations with larger charges are smaller than cations with smaller charges (e.g., V2+ has an ionic radius of 79 pm, while that of V3+
is 64 pm). Proceeding down the groups of the periodic table, we find that cations of successive elements with the same charge
generally have larger radii, corresponding to an increase in the principal quantum number, n.
An anion (negative ion) is formed by the addition of one or more electrons to the valence shell of an atom. This results in a greater
repulsion among the electrons and a decrease in Z ef f per electron. Both effects (the increased number of electrons and the
decreased Zeff) cause the radius of an anion to be larger than that of the parent atom ( Figure 3.2.3). For example, a sulfur atom
([Ne]3s23p4) has a covalent radius of 104 pm, whereas the ionic radius of the sulfide anion ([Ne]3s23p6) is 170 pm. For consecutive
elements proceeding down any group, anions have larger principal quantum numbers and, thus, larger radii.
Atoms and ions that have the same electron configuration are said to be isoelectronic. Examples of isoelectronic species are N3–,
O2–, F–, Ne, Na+, Mg2+, and Al3+ (1s22s22p6). Another isoelectronic series is P3–, S2–, Cl–, Ar, K+, Ca2+, and Sc3+ ([Ne]3s23p6). For
atoms or ions that are isoelectronic, the number of protons determines the size. The greater the nuclear charge, the smaller the
radius in a series of isoelectronic ions and atoms.
3.2.5 https://chem.libretexts.org/@go/page/119833
Figure 3.2.4 : Comparison of Ionic and Atomic Radius. Image Credit: By Popnose [CC BY-SA 3.0
(http://creativecommons.org/licenses/by-sa/3.0) or GFDL (http://www.gnu.org/copyleft/fdl.html)], via Wikimedia Commons
Summary
Glossary
covalent radius
one-half the distance between the nuclei of two identical atoms when they are joined by a covalent bond
isoelectronic
group of ions or atoms that have identical electron configurations
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Fuse School, Open Educational Resource free of charge, under a Creative Commons License: Attribution-NonCommercial CC
BY-NC (View License Deed: https://creativecommons.org/licenses/by-nc/4.0/)
Guillotined Chemistry is a production of Mark Anticole and available for free on Youtube.
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
3.2: Trends in Size is shared under a CC BY-SA license and was authored, remixed, and/or curated by LibreTexts.
3.2.6 https://chem.libretexts.org/@go/page/119833
3.2: Trends in Size (Problems)
PROBLEM 3.2.1
Based on their positions in the periodic table, predict which has the smallest atomic radius: Mg, Sr, Si, Cl, I.
Answer
Cl
PROBLEM 3.2.2
Based on their positions in the periodic table, predict which has the largest atomic radius: Li, Rb, N, F, I.
Answer
Rb
PROBLEM 3.2.3
Based on their positions in the periodic table, list the following atoms in order of increasing radius: Cs, Ca, Rb, Mg
Answer
Mg < Ca < Rb < Cs
PROBLEM 3.2.4
Based on their positions in the periodic table, list the following atoms in order of increasing radius: Sr, Cl, Ca, Si.
Answer
Cl < Si < Ca < Sr
Problem 3.2.4
PROBLEM 3.2.5
Based on their positions in the periodic table, list the following ions in order of increasing radius: K+, Ca2+, Al3+, Si4+.
Answer
Si4+ < Al3+ < Ca2+ < K+
PROBLEM 3.2.6
List the following ions in order of increasing radius: Br–, Li+, Te2–, Mg2+.
3.2.1 https://chem.libretexts.org/@go/page/119834
Answer
Li+ < Mg2+ < Br– < Te2–
Problem 3.2.6
PROBLEM 3.2.7
Which atom and/or ion is (are) isoelectronic with Br+: Se2+, Se, As–, Kr, Ga3+, Cl–?
Answer
Se, As−
PROBLEM 3.2.8
Which of the following atoms and ions is (are) isoelectronic with S2+: Si4+, Cl3+, Ar, As3+, Si, Al3+?
Answer
Si, Cl3+
Problem 3.2.8
PROBLEM 3.2.9
Compare both the numbers of protons and electrons present in each to rank the following ions in order of increasing radius:
As3–, Br–, K+, Mg2+.
Answer
Mg2+ < K+ < Br– < As3–
3.2.2 https://chem.libretexts.org/@go/page/119834
PROBLEM 3.2.10
The ionic radii of the ions S2–, Cl–, and K+ are 184, 181, 138 pm respectively. Explain why these ions have different sizes even
though they contain the same number of electrons.
Answer
They have different numbers of protons, which determines their size.
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Feedback
Think one of the answers above is wrong? Let us know here.
3.2: Trends in Size (Problems) is shared under a CC BY-SA license and was authored, remixed, and/or curated by LibreTexts.
3.2.3 https://chem.libretexts.org/@go/page/119834
3.3: Trends in Ionization Energy
Skills to Develop
Describe and explain the observed trends in ionization energy of the elements
The energy required to remove the second most loosely bound electron is called the second ionization energy (IE2).
+ 2+ −
X (g) ⟶ X (g) + e IE (3.3.2)
2
The energy required to remove the third electron is the third ionization energy, and so on. Energy is always required to remove
electrons from atoms or ions, so ionization processes are endothermic and IE values are always positive. For larger atoms, the most
loosely bound electron is located farther from the nucleus and so is easier to remove. Thus, as size (atomic radius) increases, the
ionization energy should decrease. Relating this logic to what we have just learned about radii, we would expect first ionization
energies to decrease down a group and to increase across a period.
Figure 3.3.1 : The first ionization energy of the elements in the first five periods are plotted against their atomic number.
3.3.1 https://chem.libretexts.org/@go/page/119835
Figure 3.3.1 graphs the relationship between the first ionization energy and the atomic number of several elements. Within a
period, the values of first ionization energy for the elements (IE1) generally increases with increasing Z. Down a group, the IE1
value generally decreases with increasing Z. There are some systematic deviations from this trend, however. Note that the
ionization energy of boron (atomic number 5) is less than that of beryllium (atomic number 4) even though the nuclear charge of
boron is greater by one proton. This can be explained because the energy of the subshells increases as l increases, due to
penetration and shielding (as discussed previously in this chapter). Within any one shell, the s electrons are lower in energy than the
p electrons. This means that an s electron is harder to remove from an atom than a p electron in the same shell. The electron
removed during the ionization of beryllium ([He]2s2) is an s electron, whereas the electron removed during the ionization of boron
([He]2s22p1) is a p electron; this results in a lower first ionization energy for boron, even though its nuclear charge is greater by one
proton. Thus, we see a small deviation from the predicted trend occurring each time a new subshell begins.
Figure 3.3.2 : This version of the periodic table shows the first ionization energy of (IE: 1), in kJ/mol, of selected elements.
Another deviation occurs as orbitals become more than one-half filled. The first ionization energy for oxygen is slightly less than
that for nitrogen, despite the trend in increasing IE1 values across a period. Looking at the orbital diagram of oxygen, we can see
that removing one electron will eliminate the electron–electron repulsion caused by pairing the electrons in the 2p orbital and will
result in a half-filled orbital (which is energetically favorable). Analogous changes occur in succeeding periods (note the dip for
sulfur after phosphorus in Figure 3.3.2.
Removing an electron from a cation is more difficult than removing an electron from a neutral atom because of the greater
electrostatic attraction to the cation. Likewise, removing an electron from a cation with a higher positive charge is more difficult
than removing an electron from an ion with a lower charge. Thus, successive ionization energies for one element always increase.
As seen in Table 3.3.1, there is a large increase in the ionization energies (color change) for each element. This jump corresponds
to removal of the core electrons, which are harder to remove than the valence electrons. For example, Sc and Ga both have three
valence electrons, so the rapid increase in ionization energy occurs after the third ionization.
Table 3.3.1: Successive Ionization Energies for Selected Elements (kJ/mol)
Element IE1 IE2 IE3 IE4 IE5 IE6 IE7
3.3.2 https://chem.libretexts.org/@go/page/119835
Example 3.3.1 : Ranking Ionization Energies
Predict the order of increasing energy for the following processes: IE1 for Al, IE1 for Tl, IE2 for Na, IE3 for Al.
Solution
Removing the 6p1 electron from Tl is easier than removing the 3p1 electron from Al because the higher n orbital is farther from
the nucleus, so IE1(Tl) < IE1(Al). Ionizing the third electron from
2+ 3+ −
Al (Al ⟶ Al +e )
requires more energy because the cation Al2+ exerts a stronger pull on the electron than the neutral Al atom, so IE1(Al) <
IE3(Al). The second ionization energy for sodium removes a core electron, which is a much higher energy process than
removing valence electrons. Putting this all together, we obtain:
IE1(Tl) < IE1(Al) < IE3(Al) < IE2(Na).
Exercise 3.3.1
Which has the lowest value for IE1: O, Po, Pb, or Ba?
Answer
Ba
Summary
Glossary
ionization energy
energy required to remove an electron from a gaseous atom or ion. The associated number (e.g., second ionization energy)
corresponds to the charge of the ion produced (X2+)
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
3.3.3 https://chem.libretexts.org/@go/page/119835
Guillotined Chemistry is a production of Mark Anticole and available for free on Youtube.
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
3.3: Trends in Ionization Energy is shared under a CC BY license and was authored, remixed, and/or curated by LibreTexts.
3.3.4 https://chem.libretexts.org/@go/page/119835
3.3: Trends in Ionization Energy (Problems)
PROBLEM 3.3.1
Based on their positions in the periodic table, predict which has the largest first ionization energy: Mg, Ba, B, O, Te.
Answer
O
PROBLEM 3.3.2
Based on their positions in the periodic table, predict which has the smallest first ionization energy: Li, Cs, N, F, I.
Answer
Cs
PROBLEM 3.3.3
Based on their positions in the periodic table, rank the following atoms in order of increasing first ionization energy: F, Li, N,
Rb.
Answer
Rb < Li < N < F
PROBLEM 3.3.4
Based on their positions in the periodic table, rank the following atoms or compounds in order of increasing first ionization
energy: Mg, O, S, Si
Answer
Mg< Si < S < O
PROBLEM 3.3.5
Which main group atom would be expected to have the lowest second ionization energy?
Answer
Ra
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Feedback
Think one of the answers above is wrong? Let us know here.
3.3: Trends in Ionization Energy (Problems) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.
3.3.1 https://chem.libretexts.org/@go/page/119836
3.4: Trends in Electron Affinity and Metallic Character
Skills to Develop
Describe and explain the observed trends in electron affinity and metallic character of the elements
To recognize the inverse relationship of ionization energies and electron affinities
This process can be either endothermic or exothermic, depending on the element. The EA of some of the elements is given in
Figure 3.4.6. You can see that many of these elements have negative values of EA, which means that energy is released when the
gaseous atom accepts an electron. However, for some elements, energy is required for the atom to become negatively charged and
the value of their EA is positive. Just as with ionization energy, subsequent EA values are associated with forming ions with more
charge. The second EA is the energy associated with adding an electron to an anion to form a –2 ion, and so on.
Figure 3.4.1 : This version of the periodic table displays the electron affinity values (in kJ/mol) for selected elements.:
As we might predict, it becomes easier to add an electron across a series of atoms as the effective nuclear charge of the atoms
increases. We find, as we go from left to right across a period, EAs tend to become more negative. The exceptions found among the
elements of group 2 (2A), group 15 (5A), and group 18 (8A) can be understood based on the electronic structure of these groups.
The noble gases, group 18 (8A), have a completely filled shell and the incoming electron must be added to a higher n level, which
3.4.1 https://chem.libretexts.org/@go/page/119837
is more difficult to do. Group 2 (2A) has a filled ns subshell, and so the next electron added goes into the higher energy np, so,
again, the observed EA value is not as the trend would predict. Finally, group 15 (5A) has a half-filled np subshell and the next
electron must be paired with an existing np electron. In all of these cases, the initial relative stability of the electron configuration
disrupts the trend in EA.
We also might expect the atom at the top of each group to have the largest EA; their first ionization potentials suggest that these
atoms have the largest effective nuclear charges. However, as we move down a group, we see that the second element in the group
most often has the greatest EA. The reduction of the EA of the first member can be attributed to the small size of the n = 2 shell and
the resulting large electron–electron repulsions. For example, chlorine, with an EA value of –348 kJ/mol, has the highest value of
any element in the periodic table. The EA of fluorine is –322 kJ/mol. When we add an electron to a fluorine atom to form a fluoride
anion (F–), we add an electron to the n = 2 shell. The electron is attracted to the nucleus, but there is also significant repulsion from
the other electrons already present in this small valence shell. The chlorine atom has the same electron configuration in the valence
shell, but because the entering electron is going into the n = 3 shell, it occupies a considerably larger region of space and the
electron–electron repulsions are reduced. The entering electron does not experience as much repulsion and the chlorine atom
accepts an additional electron more readily.
The properties discussed in this unit (size of atoms and ions, effective nuclear charge, ionization energies, and electron affinities)
are central to understanding chemical reactivity. For example, because fluorine has an energetically favorable EA and a large
energy barrier to ionization (IE), it is much easier to form fluorine anions than cations.
3.4.2 https://chem.libretexts.org/@go/page/119837
Figure 3.4.2 : Metallic character increases as you go from right to left on periods and move down groups on the periodic table.
The gray squares represent metals, the orange squares are "metalloids" and the yellow squares are nonmetals.
Credit: "Various Periodic Trends" by Sandbh is licensed under CC BY-SA 4.0
Metallic properties including conductivity and malleability (the ability to be formed into sheets) depend on having electrons that
can be removed easily. Thus, metallic character increases as we move down a group and decreases across a period in the same trend
observed for atomic size because it is easier to remove an electron that is farther away from the nucleus.
Summary
Electron affinity (the energy associated with forming an anion) is more favorable (exothermic) when electrons are placed into
lower energy orbitals, closer to the nucleus. Therefore, electron affinity becomes increasingly negative as we move left to right
across the periodic table and decreases as we move down a group. For both IE and electron affinity data, there are exceptions to the
trends when dealing with completely filled or half-filled subshells. Metallic character increases from right to left across periods on
the periodic table and down columns.
Glossary
electron affinity
energy required to add an electron to a gaseous atom to form an anion
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
Guillotined Chemistry is a production of Mark Anticole and available for free on Youtube.
3.4.3 https://chem.libretexts.org/@go/page/119837
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
3.4: Trends in Electron Affinity and Metallic Character is shared under a CC BY license and was authored, remixed, and/or curated by
LibreTexts.
3.4.4 https://chem.libretexts.org/@go/page/119837
3.4: Trends in Electron Affinity and Metallic Character (Problems)
PROBLEM 3.4.1
Based on their positions in the periodic table, predict which has the largest electron affinity: Mg, Ba, B, O, Te.
Answer
O
PROBLEM 3.4.2
Based on their positions in the periodic table, predict which has the smallest electron affinity: Li, Cs, N, F, I.
Answer
Cs
PROBLEM 3.4.3
Based on their positions in the periodic table, rank the following atoms in order of increasing electron affinity: F, Li, N, Rb.
Answer
Rb < Li < N < F
PROBLEM 3.4.4
Based on their positions in the periodic table, rank the following atoms or compounds in order of increasing electron affnity:
Mg, O, S, Si
Answer
S < Si < Mg < O
PROBLEM 3.4.5
Based on their positions in the periodic table, rank the following atoms in order of increasing metallic character: F, Li, N, Rb.
Answer
F < N < Li < Rb
PROBLEM 3.4.6
Based on their positions in the periodic table, predict which is the most metallic element: Mg, Ba, B, O, Te.
Answer
Ba
PROBLEM 3.4.7
Based on their positions in the periodic table, predict which is the most nonmetallic element: Li, Cs, N, F, I.
Answer
F
3.4.1 https://chem.libretexts.org/@go/page/119838
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Feedback
Think one of the answers above is wrong? Let us know here.
3.4: Trends in Electron Affinity and Metallic Character (Problems) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed,
and/or curated by LibreTexts.
3.4.2 https://chem.libretexts.org/@go/page/119838
CHAPTER OVERVIEW
Unit 4: Lewis Structures
Unit Objectives
Write Lewis symbols for neutral atoms and ions
Draw Lewis structures depicting the bonding in simple molecules
Compute formal charges for atoms in any Lewis structure
Use formal charges to identify the most reasonable Lewis structure for a given molecule
Identify the oxidation states of atoms in Lewis structures
Unit 4: Lewis Structures is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
1
4.1: Lewis Dot Diagrams
Skills to Develop
Write Lewis symbols for neutral atoms and ions
Thus far, we have discussed the various types of bonds that form between atoms and/or ions. In all cases, these bonds involve the
sharing or transfer of valence shell electrons between atoms. In this section, we will explore the typical method for depicting
valence shell electrons and chemical bonds, namely Lewis symbols and Lewis structures.
Lewis Symbols
We use Lewis symbols to describe valence electron configurations of atoms and monatomic ions. A Lewis symbol consists of an
elemental symbol surrounded by one dot for each of its valence electrons:
Figure 4.1.1 shows the Lewis symbols for the elements of the third period of the periodic table. Electron dots are typically arranged
in four pairs located on the four "sides" of the atomic symbol.
Figure 4.1.1 : Lewis symbols illustrating the number of valence electrons for each element in the third period of the periodic table.
Lewis symbols can be used to illustrate the formation of cations from atoms, as shown here for sodium and calcium:
Likewise, they can be used to show the formation of anions from atoms, as shown here for chlorine and sulfur:
Figure 4.1.2 demonstrates the use of Lewis symbols to show the transfer of electrons during the formation of ionic compounds.
4.1.1 https://chem.libretexts.org/@go/page/119803
Figure 4.1.2 : Cations are formed when atoms lose electrons, represented by fewer Lewis dots, whereas anions are formed by
atoms gaining electrons. The total number of electrons does not change.
Summary
Valence electronic structures can be visualized by drawing Lewis symbols (for atoms and monatomic ions) . Lone pairs, unpaired
electrons, and single, double, or triple bonds are used to indicate where the valence electrons are located around the atom.
Glossary
Lewis symbol
symbol for an element or monatomic ion that uses a dot to represent each valence electron in the element or ion
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
4.1: Lewis Dot Diagrams is shared under a CC BY license and was authored, remixed, and/or curated by LibreTexts.
4.1.2 https://chem.libretexts.org/@go/page/119803
4.1: Lewis Dot Diagrams (Problems)
PROBLEM 4.1.1
Write the Lewis symbols for each of the following ions:
a. As3–
b. I–
c. Be2+
d. O2–
e. Ga3+
f. Li+
g. N3–
Answer a
eight electrons:
Answer b
eight electrons:
Answer c
no electrons
Be2+
Answer d
eight electrons:
Answer e
no electrons
Ga3+
Answer f
no electrons
Li+
Answer g
eight electrons:
PROBLEM 4.1.2
4.1.1 https://chem.libretexts.org/@go/page/119804
Write the Lewis symbols for the following elements:
a. Cl
b. Na
c. Mg
d. Ca
e. K
f. Br
g. Sr
h. F
Answer a
Answer b
Answer c
Answer d
Answer e
Answer f
Answer g
4.1.2 https://chem.libretexts.org/@go/page/119804
Answer h
PROBLEM 4.1.3
Write the Lewis symbols of the ions in each of the following ionic compounds and the Lewis symbols of the atom from which
they are formed:
a. MgS
b. Al2O3
c. GaCl3
d. K2O
e. Li3N
f. KF
Answer a
Two Lewis structures are shown. The left shows the symbol M g with a superscripted two positive sign while the right shows the symbol S surrounded by eight dots and a superscripted two negative sign.
Answer b
Two Lewis structures are shown. The left shows the symbol A l with a superscripted three positive sign while the right shows the symbol O surrounded by eight dots and a superscripted two negative sign.
Answer c
Two Lewis structures are shown. The left shows the symbol G a with a superscripted three positive sign while the right shows the symbol C l surrounded by eight dots and a superscripted negative sign.
Answer d
Two Lewis structures are shown. The left shows the symbol K with a superscripted positive sign while the right shows the symbol O surrounded by eight dots and a superscripted two negative sign.
Answer e
Two Lewis structures are shown. The left shows the symbol L i with a superscripted positive sign while the right shows the symbol N surrounded by eight dots and a superscripted three negative sign.
Answer f
Two Lewis structures are shown. The left shows the symbol K with a superscripted positive sign while the right shows the symbol F surrounded by eight dots and a superscripted negative sign.
PROBLEM 4.1.4
In the Lewis structures listed here, M and X represent various elements in the third period of the periodic table. Write the
formula of each compound using the chemical symbols of each element:
(a)
Two Lewis structures are shown side-by-side, each surrounded by brackets. The left structure shows the symbol M with a superscripted two positive sign. The right shows the symbol X surrounded by four lone
pairs of electrons with a superscripted two negative sign outside of the brackets.
(b)
Two Lewis structures are shown side-by-side, each surrounded by brackets. The left structure shows the symbol M with a superscripted three positive sign. The right structure shows the symbol X surrounded by
four lone pairs of electrons with a superscripted negative sign and a subscripted three both outside of the brackets.
(c)
4.1.3 https://chem.libretexts.org/@go/page/119804
Two Lewis structures are shown side-by-side, each surrounded by brackets. The left structure shows the symbol M with a superscripted positive sign and a subscripted two outside of the brackets. The right
structure shows the symbol X surrounded by four lone pairs of electrons with a superscripted two negative sign outside of the brackets.
(d)
Two Lewis structures are shown side-by-side, each surrounded by brackets. The left structure shows the symbol M with a superscripted three positive sign and a subscripted two outside of the brackets. The right
structure shows the symbol X surrounded by four lone pairs of electrons with a superscripted two negative sign and subscripted three both outside of the brackets.
Answer a
MgS
Answer b
AlCl3
Answer c
Na2S
Answer d
Al2S3
PROBLEM 4.1.5
Identify the atoms that correspond to each of the following electron configurations.
a. 1s22s22p5
b. 1s22s22p63s2
c. 1s22s22p63s23p64s23d104p4
d. 1s22s22p63s23p64s23d104p1
Answer a
F
Answer b
Mg
Answer c
Se
Answer d
Ga
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
4.1.4 https://chem.libretexts.org/@go/page/119804
Feedback
Think one of the answers above is wrong? Let us know here.
4.1: Lewis Dot Diagrams (Problems) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
4.1.5 https://chem.libretexts.org/@go/page/119804
4.2: Lewis Structures
Skills to Develop
Draw Lewis structures depicting the bonding in simple molecules
Lewis Structures
We also use Lewis symbols to indicate the formation of covalent bonds, which are shown in Lewis structures, drawings that
describe the bonding in molecules and polyatomic ions. For example, when two chlorine atoms form a chlorine molecule, they
share one pair of electrons:
The Lewis structure indicates that each Cl atom has three pairs of electrons that are not used in bonding (called lone pairs) and one
shared pair of electrons (written between the atoms). A dash (or line) is usually used to indicate a shared pair of electrons:
In the Lewis model, a single shared pair of electrons is a single bond. Each Cl atom interacts with eight valence electrons total: the
six in the lone pairs and the two in the single bond.
Group 15 elements such as nitrogen have five valence electrons in the atomic Lewis symbol: one lone pair and three unpaired
electrons. To obtain an octet, these atoms form three covalent bonds, as in NH3 (ammonia). Oxygen and other atoms in group 16
obtain an octet by forming two covalent bonds:
4.2.1 https://chem.libretexts.org/@go/page/119881
electrons are shared between a pair of atoms, as between the carbon and oxygen atoms in CH2O (formaldehyde) and between the
two carbon atoms in C2H4 (ethylene):
A triple bond forms when three electron pairs are shared by a pair of atoms, as in carbon monoxide (CO) and the cyanide ion
(CN–):
For more complicated molecules and molecular ions, it is helpful to follow the step-by-step procedure outlined here:
1. Determine the total number of valence (outer shell) electrons among all the atoms. For cations, subtract one electron for each
positive charge. For anions, add one electron for each negative charge.
2. Draw a skeleton structure of the molecule or ion, arranging the atoms around a central atom. (Generally, the least
electronegative element should be placed in the center.) Connect each atom to the central atom with a single bond (one electron
pair).
3. Distribute the remaining electrons as lone pairs on the terminal atoms (except hydrogen), completing an octet around each
atom.
4. Place all remaining electrons on the central atom.
5. Rearrange the electrons of the outer atoms to make multiple bonds with the central atom in order to obtain octets wherever
possible.
Let us determine the Lewis structures of SiH4, CHO , NO+, and OF2 as examples in following this procedure:
−
2
1. Determine the total number of valence (outer shell) electrons in the molecule or ion.
For a molecule, we add the number of valence electrons on each atom in the molecule:
SiH (4.2.1)
4
For a negative ion, such as CHO , we add the number of valence electrons on the atoms to the number of negative charges
−
on the ion (one electron is gained for each single negative charge):
4.2.2 https://chem.libretexts.org/@go/page/119881
−
CHO
2
+ 1 additional electron = 1
––––––––––––––––––––––––––––––––––––––––
= 18 valence electrons
For a positive ion, such as NO+, we add the number of valence electrons on the atoms in the ion and then subtract the
number of positive charges on the ion (one electron is lost for each single positive charge) from the total number of valence
electrons:
+
NO
2. Draw a skeleton structure of the molecule or ion, arranging the atoms around a central atom and connecting each atom to the
central atom with a single (one electron pair) bond. (Note that we denote ions with brackets around the structure, indicating the
charge outside the brackets:)
3. When several arrangements of atoms are possible, as for CHO , we must use experimental evidence to choose the correct one.
−
In general, the less electronegative elements are more likely to be central atoms. In CHO , the less electronegative carbon
−
atom occupies the central position with the oxygen and hydrogen atoms surrounding it. Other examples include P in POCl3, S in
SO2, and Cl in ClO . An exception is that hydrogen is almost never a central atom. As the most electronegative element,
−
For OF2, we had 16 electrons remaining in Step 3, and we placed 12, leaving 4 to be placed on the central atom:
6. Rearrange the electrons of the outer atoms to make multiple bonds with the central atom in order to obtain octets wherever
possible.
SiH4: Si already has an octet, so nothing needs to be done.
CHO : We have distributed the valence electrons as lone pairs on the oxygen atoms, but the carbon atom lacks an octet:
−
4.2.3 https://chem.libretexts.org/@go/page/119881
NO+: For this ion, we added eight outer electrons, but neither atom has an octet. We cannot add any more electrons since we
have already used the total that we found in Step 1, so we must move electrons to form a multiple bond:
This still does not produce an octet, so we must move another pair, forming a triple bond:
4.2.4 https://chem.libretexts.org/@go/page/119881
NH3: two electrons placed on nitrogen
Where needed, rearrange electrons to form multiple bonds in order to obtain an octet on each atom:
HCN: form two more C–N bonds
H3CCH3: all atoms have the correct number of electrons
HCCH: form a triple bond between the two carbon atoms
NH3: all atoms have the correct number of electrons
Exercise 4.2.1
Both carbon monoxide, CO, and carbon dioxide, CO2, are products of the combustion of fossil fuels. Both of these gases also
cause problems: CO is toxic and CO2 has been implicated in global climate change. What are the Lewis structures of these two
molecules?
Answer
Fullerene Chemistry
Carbon soot has been known to man since prehistoric times, but it was not until fairly recently that the molecular structure of the
main component of soot was discovered. In 1996, the Nobel Prize in Chemistry was awarded to Richard Smalley, Robert Curl,
and Harold Kroto for their work in discovering a new form of carbon, the C60 buckminsterfullerene molecule. An entire class of
compounds, including spheres and tubes of various shapes, were discovered based on C60. This type of molecule, called a
fullerene, consists of a complex network of single- and double-bonded carbon atoms arranged in such a way that each carbon
atom obtains a full octet of electrons. Because of their size and shape, fullerenes can encapsulate other molecules, so they have
shown potential in various applications from hydrogen storage to targeted drug delivery systems. They also possess unique
electronic and optical properties that have been put to good use in solar powered devices and chemical sensors.
Odd-electron Molecules
We call molecules that contain an odd number of electrons free radicals. Nitric oxide, NO, is an example of an odd-electron
molecule; it is produced in internal combustion engines when oxygen and nitrogen react at high temperatures.
To draw the Lewis structure for an odd-electron molecule like NO, we follow the same five steps we would for other molecules,
but with a few minor changes:
1. Determine the total number of valence (outer shell) electrons. The sum of the valence electrons is 5 (from N) + 6 (from O) = 11.
The odd number immediately tells us that we have a free radical, so we know that not every atom can have eight electrons in its
valence shell.
2. Draw a skeleton structure of the molecule. We can easily draw a skeleton with an N–O single bond: N–O
3. Distribute the remaining electrons as lone pairs on the terminal atoms. In this case, there is no central atom, so we distribute the
electrons around both atoms. We give eight electrons to the more electronegative atom in these situations; thus oxygen has the
4.2.5 https://chem.libretexts.org/@go/page/119881
filled valence shell:
4. Place all remaining electrons on the central atom. Since there are no remaining electrons, this step does not apply.
5. Rearrange the electrons to make multiple bonds with the central atom in order to obtain octets wherever possible. We know that
an odd-electron molecule cannot have an octet for every atom, but we want to get each atom as close to an octet as possible. In
this case, nitrogen has only five electrons around it. To move closer to an octet for nitrogen, we take one of the lone pairs from
oxygen and use it to form a NO double bond. (We cannot take another lone pair of electrons on oxygen and form a triple bond
because nitrogen would then have nine electrons:)
Electron-deficient Molecules
We will also encounter a few molecules that contain central atoms that do not have a filled valence shell. Generally, these are
molecules with central atoms from groups 2 and 13 and outer atoms that are hydrogen or other atoms that do not form multiple
bonds. For example, in the Lewis structures of beryllium dihydride, BeH2, and boron trifluoride, BF3, the beryllium and boron
atoms each have only four and six electrons, respectively. It is possible to draw a structure with a double bond between a boron
atom and a fluorine atom in BF3, satisfying the octet rule, but experimental evidence indicates the bond lengths are closer to that
expected for B–F single bonds. This suggests the best Lewis structure has three B–F single bonds and an electron deficient boron.
The reactivity of the compound is also consistent with an electron deficient boron. However, the B–F bonds are slightly shorter
than what is actually expected for B–F single bonds, indicating that some double bond character is found in the actual molecule.
An atom like the boron atom in BF3, which does not have eight electrons, is very reactive. It readily combines with a molecule
containing an atom with a lone pair of electrons. For example, NH3 reacts with BF3 because the lone pair on nitrogen can be shared
with the boron atom:
Hypervalent Molecules
Elements in the second period of the periodic table (n = 2) can accommodate only eight electrons in their valence shell orbitals
because they have only four valence orbitals (one 2s and three 2p orbitals). Elements in the third and higher periods (n ≥ 3) have
more than four valence orbitals and can share more than four pairs of electrons with other atoms because they have empty d orbitals
in the same shell. Molecules formed from these elements are sometimes called hypervalent molecules.Table 4.2.5 shows the Lewis
structures for two hypervalent molecules, PCl5 and SF6.
Table 4.2.5 : In PCl5, the central atom phosphorus shares five pairs of electrons. In SF6, sulfur shares six pairs of electrons.
In some hypervalent molecules, such as IF5 and XeF4, some of the electrons in the outer shell of the central atom are lone pairs:
4.2.6 https://chem.libretexts.org/@go/page/119881
When we write the Lewis structures for these molecules, we find that we have electrons left over after filling the valence shells of
the outer atoms with eight electrons. These additional electrons must be assigned to the central atom.
Solution
We can draw the Lewis structure of any covalent molecule by following the six steps discussed earlier. In this case, we can
condense the last few steps, since not all of them apply.
Step 1: Calculate the number of valence electrons:
XeF
2
: 8 + (2 × 7) = 22
XeF
6
: 8 + (6 × 7) = 50
Step 2: Draw a skeleton joining the atoms by single bonds. Xenon will be the central atom because fluorine cannot be a
central atom:
XeF6: We place three lone pairs of electrons around each F atom, accounting for 36 electrons. Two electrons remain, and this
lone pair is placed on the Xe atom:
Answer
4.2.7 https://chem.libretexts.org/@go/page/119881
Summary
Video 4.2.1 : Need more help with Lewis Structures? This video provides a great explanation!
Valence electronic structures can be visualized by drawing Lewis symbols (for atoms and monatomic ions) and Lewis structures
(for molecules and polyatomic ions). Lone pairs, unpaired electrons, and single, double, or triple bonds are used to indicate where
the valence electrons are located around each atom in a Lewis structure. Most structures—especially those containing second row
elements—obey the octet rule, in which every atom (except H) is surrounded by eight electrons. Exceptions to the octet rule occur
for odd-electron molecules (free radicals), electron-deficient molecules, and hypervalent molecules.
Glossary
double bond
covalent bond in which two pairs of electrons are shared between two atoms
free radical
molecule that contains an odd number of electrons
hypervalent molecule
molecule containing at least one main group element that has more than eight electrons in its valence shell
Lewis structure
diagram showing lone pairs and bonding pairs of electrons in a molecule or an ion
Lewis symbol
symbol for an element or monatomic ion that uses a dot to represent each valence electron in the element or ion
lone pair
two (a pair of) valence electrons that are not used to form a covalent bond
octet rule
guideline that states main group atoms will form structures in which eight valence electrons interact with each nucleus, counting
bonding electrons as interacting with both atoms connected by the bond
single bond
4.2.8 https://chem.libretexts.org/@go/page/119881
bond in which a single pair of electrons is shared between two atoms
triple bond
bond in which three pairs of electrons are shared between two atoms
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
4.2: Lewis Structures is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
4.2.9 https://chem.libretexts.org/@go/page/119881
4.2: Lewis Structures (Problems)
PROBLEM 4.2.1
Write Lewis structures for the following: (please note, none of the solutions are using the expanded octet rule or formal charges)
a. H2
b. HBr
c. PCl3
d. SF2
e. H2CCH2
f. HNNH
g. H2CNH
h. NO–
i. N2
j. CO
k. CN–
Answer a
Answer b
Answer c
Answer d
Answer e
4.2.1 https://chem.libretexts.org/@go/page/125538
Answer f
Answer g
Answer h
Answer i
Answer j
4.2.2 https://chem.libretexts.org/@go/page/125538
Answer k
PROBLEM 4.2.2
Write Lewis structures for the following: (please note, none of the solutions are using the expanded octet rule or formal charges)
a. O2
b. H2CO
c. ClNO
d. SiCl4
e. H3O+
f. NH +
g. BF −
h. HCCH
i. ClCN
j. C 2+
2
Answer a
A Lewis structure shows two oxygen atoms double bonded together, and each has two lone pairs of electrons.
Answer b
A Lewis structure shows a carbon atom that is single bonded to two hydrogen atoms and double bonded to an oxygen atom. The oxygen atom has two lone pairs of electrons.
Answer c
A Lewis structure shows a nitrogen atom with a lone pair of electrons single bonded to a chlorine atom that has three lone pairs of electrons. The nitrogen is also double bonded to an oxygen which has two
lone pairs of electrons.
Answer d
A Lewis structure shows a silicon atom that is single bonded to four chlorine atoms. Each chlorine atom has three lone pairs of electrons.
Answer e
A Lewis structure shows an oxygen atom with a lone pair of electrons single bonded to three hydrogen atoms. The structure is surrounded by brackets with a superscripted positive sign.
Answer f
A Lewis structure shows a nitrogen atom single bonded to four hydrogen atoms. The structure is surrounded by brackets with a superscripted positive sign.
Answer g
A Lewis structure shows a boron atom single bonded to four fluorine atoms. Each fluorine atom has three lone pairs of electrons. The structure is surrounded by brackets with a superscripted negative sign.
Answer h
A Lewis structure shows two carbon atoms that are triple bonded together. Each carbon is also single bonded to a hydrogen atom.
Answer i
A Lewis structure shows a carbon atom that is triple bonded to a nitrogen atom that has one lone pair of electrons. The carbon is also single bonded to a chlorine atom that has three lone pairs of electrons.
4.2.3 https://chem.libretexts.org/@go/page/125538
Answer j
A Lewis structure shows two carbon atoms joined with a triple bond. A superscripted 2 positive sign lies to the right of the second carbon.
PROBLEM 4.2.3
Write Lewis structures for: (please note, none of the solutions are using the expanded octet rule or formal charges)
a. PO 3−
c. SO2−
3
d. HONO
Answer a
Answer b
Answer c
4.2.4 https://chem.libretexts.org/@go/page/125538
Problem 4.3.3
PROBLEM 4.2.4
Methanol, H3COH, is used as the fuel in some race cars. Ethanol, C2H5OH, is used extensively as motor fuel in Brazil. Both
methanol and ethanol produce CO2 and H2O when they burn. Write the chemical equations for these combustion reactions using
Lewis structures instead of chemical formulas.
Answer
Two reactions are shown using Lewis structures. The top reaction shows a carbon atom, single bonded to three hydrogen atoms and single bonded to an oxygen atom with two lone pairs of electrons. The
oxygen atom is also bonded to a hydrogen atom. This is followed by a plus sign and the number one point five, followed by two oxygen atoms bonded together with a double bond and each with two lone pairs
of electrons. A right-facing arrow leads to a carbon atom that is double bonded to two oxygen atoms, each of which has two lone pairs of electrons. This structure is followed by a plus sign, a number two, and
a structure made up of an oxygen with two lone pairs of electrons single bonded to two hydrogen atoms. The bottom reaction shows a carbon atom, single bonded to three hydrogen atoms and single bonded to
another carbon atom. The second carbon atom is single bonded to two hydrogen atoms and one oxygen atom with two lone pairs of electrons. The oxygen atom is also bonded to a hydrogen atom. This is
followed by a plus sign and the number three, followed by two oxygen atoms bonded together with a double bond. Each oxygen atom has two lone pairs of electrons. A right-facing arrow leads to a number
two and a carbon atom that is double bonded to two oxygen atoms, each of which has two lone pairs of electrons. This structure is followed by a plus sign, a number three, and a structure made up of an
oxygen with two lone pairs of electrons single bonded to two hydrogen atoms.
PROBLEM 4.2.5
Many planets in our solar system contain organic chemicals including methane (CH4) and traces of ethylene (C2H4), ethane
(C2H6), propyne (H3CCCH), and diacetylene (HCCCCH). Write the Lewis structures for each of these molecules. (diacetylene
may be a little tricky!)
Answer
4.2.5 https://chem.libretexts.org/@go/page/125538
Problem 4.3.5
PROBLEM 4.2.6
Carbon tetrachloride was formerly used in fire extinguishers for electrical fires. It is no longer used for this purpose because of
the formation of the toxic gas phosgene, Cl2CO. Write the Lewis structures for carbon tetrachloride and phosgene.
Answer
Two Lewis structures are shown. The left depicts a carbon atom single bonded to four chlorine atoms, each with three lone pairs of electrons. The right shows a carbon atom double bonded to an oxygen
atom that has two lone pairs of electrons. The carbon atom is also single bonded to two chlorine atoms, each of which has three lone pairs of electrons.
PROBLEM 4.2.7
The arrangement of atoms in several biologically important molecules is given here. Complete the Lewis structures of these
molecules by adding multiple bonds and lone pairs. Do not add any more atoms.
a. the amino acid serine:
A Lewis structure is shown. A nitrogen atom is single bonded to two hydrogen atoms and a carbon atom. The carbon atom is single bonded to a hydrogen atom and two other carbon atoms. One of these carbon
atoms is single bonded to two hydrogen atoms and an oxygen atom. The oxygen atom is bonded to a hydrogen atom. The other carbon atom is single bonded to two oxygen atoms, one of which is bonded to a
hydrogen atom.
b. urea:
A Lewis structure is shown. A nitrogen atom is single bonded to two hydrogen atoms and a carbon atom. The carbon atom is single bonded to an oxygen atom and another nitrogen atom. That nitrogen atom is
then single bonded to two hydrogen atoms.
c. pyruvic acid:
A Lewis structure is shown. A carbon atom is single bonded to three hydrogen atoms and another carbon atom. The second carbon atom is single bonded to an oxygen atom and a third carbon atom. This carbon
is then single bonded to two oxygen atoms, one of which is single bonded to a hydrogen atom.
d. uracil:
A Lewis hexagonal ring structure is shown. From the top of the ring (moving clockwise), three carbon atoms, one nitrogen atom, a carbon atom, and a nitrogen atom are single bonded to each another. The top
carbon atom is single bonded to an oxygen atom. The second and third carbons and the nitrogen atom are each single bonded to a hydrogen atom. The next carbon atom is single bonded to an oxygen atom, and the
last nitrogen atom is single bonded to a hydrogen atom.
e. carbonic acid:
A Lewis structure is shown. A carbon atom is single bonded to three oxygen atoms. Two of those oxygen atoms are each single bonded to a hydrogen atom.
Answer a
A Lewis structure is shown. A nitrogen atom is single bonded to two hydrogen atoms and a carbon atom. The carbon atom is single bonded to a hydrogen atom and two other carbon atoms. One of these
carbon atoms is single bonded to two hydrogen atoms and an oxygen atom. The oxygen atom is bonded to a hydrogen atom. The other carbon is single bonded to two oxygen atoms, one of which is bonded to
a hydrogen atom. The oxygen atoms have two lone pairs of electron dots, and the nitrogen atom has one lone pair of electron dots.
Answer b
A Lewis structure is shown. A nitrogen atom is single bonded to two hydrogen atoms and a carbon atom. The carbon atom is single bonded to an oxygen atom and one nitrogen atom. That nitrogen atom is
then single bonded to two hydrogen atoms. The oxygen atom has two lone pairs of electron dots, and the nitrogen atoms have one lone pair of electron dots each.
Answer c
A Lewis structure is shown. A carbon atom is single bonded to three hydrogen atoms and a carbon atom. The carbon atom is single bonded to an oxygen atom and a third carbon atom. This carbon is then
single bonded to two oxygen atoms, one of which is single bonded to a hydrogen atom. Each oxygen atom has two lone pairs of electron dots.
4.2.6 https://chem.libretexts.org/@go/page/125538
Answer d
A Lewis hexagonal ring structure is shown. From the top of the ring, three carbon atoms, one nitrogen atom, a carbon atom and a nitrogen atom are single bonded to one another. The top carbon is single
bonded to an oxygen, the second and third carbons and the nitrogen atom are each single bonded to a hydrogen atom. The next carbon is single bonded to an oxygen atom and the last nitrogen is single bonded
to a hydrogen atom. The oxygen atoms have two lone pairs of electron dots, and the nitrogen atoms have one lone pair of electron dots.
Answer e
A Lewis structure is shown. A carbon atom is single bonded to three oxygen atoms. Two of those oxygen atoms are each single bonded to a hydrogen atom. Each oxygen atom has two lone pairs of electron
dots.
PROBLEM 4.2.8
How are single, double, and triple bonds similar? How do they differ?
Answer
Each bond includes a sharing of electrons between atoms. Two electrons are shared in a single bond; four electrons are
shared in a double bond; and six electrons are shared in a triple bond.
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Feedback
Think one of the answers above is wrong? Let us know here.
4.2: Lewis Structures (Problems) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
4.2.7 https://chem.libretexts.org/@go/page/125538
4.3: Formal Charge and Oxidation State
Skills to Develop
Compute formal charges for atoms in any Lewis structure
Use formal charges to identify the most reasonable Lewis structure for a given molecule
Identify the oxidation states of atoms in Lewis structures
Previously, we discussed how to write Lewis structures for molecules and polyatomic ions. In some cases, however, there is
seemingly more than one valid structure for a molecule. We can use the concept of formal charges to help us predict the most
appropriate Lewis structure when more than one is reasonable. But first, let's introduce a concept we will refer back to frequently
for the rest of this term: electronegativity.
Electronegativity
Whether a bond is nonpolar or polar covalent is determined by a property of the bonding atoms called electronegativity.
Electronegativity is a measure of the tendency of an atom to attract electrons (or electron density) towards itself. It determines how
the shared electrons are distributed between the two atoms in a bond. The more strongly an atom attracts the electrons in its bonds,
the larger its electronegativity. Electrons in a polar covalent bond are shifted toward the more electronegative atom; thus, the more
electronegative atom is the one with the partial negative charge. The greater the difference in electronegativity, the more polarized
the electron distribution and the larger the partial charges of the atoms.
Figure 4.3.1 shows the electronegativity values of the elements as proposed by one of the most famous chemists of the twentieth
century: Linus Pauling. In general, electronegativity increases from left to right across a period in the periodic table and decreases
down a group. Thus, the nonmetals, which lie in the upper right, tend to have the highest electronegativities, with fluorine the most
electronegative element of all (EN = 4.0). Metals tend to be less electronegative elements, and the group 1 metals have the lowest
electronegativities. Note that noble gases are excluded from this figure because these atoms usually do not share electrons with
others atoms since they have a full valence shell. (While noble gas compounds such as XeO2 do exist, they can only be formed
under extreme conditions, and thus they do not fit neatly into the general model of electronegativity.)
Figure 4.3.1 : The electronegativity values derived by Pauling follow predictable periodic trends with the higher
electronegativities toward the upper right of the periodic table.
4.3.1 https://chem.libretexts.org/@go/page/119805
We can double-check formal charge calculations by determining the sum of the formal charges for the whole structure. The sum of
the formal charges of all atoms in a molecule must be zero; the sum of the formal charges in an ion should equal the charge of the
ion.
We must remember that the formal charge calculated for an atom is not the actual charge of the atom in the molecule. Formal
charge is only a useful bookkeeping procedure; it does not indicate the presence of actual charges.
Solution
We divide the bonding electron pairs equally for all I– Cl bonds:
We assign lone pairs of electrons to their atoms. Each Cl atom now has seven electrons assigned to it, and the I atom has eight.
Subtract this number from the number of valence electrons for the neutral atom:
I: 7 – 8 = –1
Cl: 7 – 7 = 0
The sum of the formal charges of all the atoms equals –1, which is identical to the charge of the ion (–1).
Exercise 4.3.1
Calculate the formal charge for each atom in the carbon monoxide molecule:
Answer
C −1, O +1
Solution
Assign one of the electrons in each Br–Cl bond to the Br atom and one to the Cl atom in that bond:
Assign the lone pairs to their atom. Now each Cl atom has seven electrons and the Br atom has seven electrons.
Subtract this number from the number of valence electrons for the neutral atom. This gives the formal charge:
Br: 7 – 7 = 0
Cl: 7 – 7 = 0
All atoms in BrCl have a formal charge of zero, and the sum of the formal charges totals zero, as it must in a neutral molecule.
3
Exercise 4.3.2
Determine the formal charge for each atom in NCl . 3
4.3.2 https://chem.libretexts.org/@go/page/119805
Answer
N: 0; all three Cl atoms: 0
To see how these guidelines apply, let us consider some possible structures for carbon dioxide, CO . We know from our previous
2
discussion that the less electronegative atom typically occupies the central position, but formal charges allow us to understand why
this occurs. We can draw three possibilities for the structure: carbon in the center and double bonds, carbon in the center with a
single and triple bond, and oxygen in the center with double bonds:
Comparing the three formal charges, we can definitively identify the structure on the left as preferable because it has only formal
charges of zero (Guideline 1).
As another example, the thiocyanate ion, an ion formed from a carbon atom, a nitrogen atom, and a sulfur atom, could have three
different molecular structures: CNS , NCS , or CSN . The formal charges present in each of these molecular structures can help
– – –
us pick the most likely arrangement of atoms. Possible Lewis structures and the formal charges for each of the three possible
structures for the thiocyanate ion are shown here:
4.3.3 https://chem.libretexts.org/@go/page/119805
Note that the sum of the formal charges in each case is equal to the charge of the ion (–1). However, the first arrangement of atoms
is preferred because it has the lowest number of atoms with nonzero formal charges (Guideline 2). Also, it places the least
electronegative atom in the center, and the negative charge on the more electronegative element (Guideline 4).
The structure with a terminal oxygen atom best satisfies the criteria for the most stable distribution of formal charge:
The number of atoms with formal charges are minimized (Guideline 2), and there is no formal charge larger than one (Guideline
2). This is again consistent with the preference for having the less electronegative atom in the central position.
Exercise 4.3.3
Which is the most likely molecular structure for the nitrite (NO ) ion?
−
Answer
–
ONO
Oxidation Numbers
We have also discussed electronegativity, which gives rise to polarity in bonds and molecules. Thus, sometimes it is helpful for us
to define another somewhat artificial device - invented by chemists, not by molecules - which enables us to keep track of electrons
in complicated reactions where electrons rearrange into new bonds.
We can obtain oxidation numbers by arbitrarily assigning the electrons of each covalent bond to the more electronegative atom in
the bond. This is in contrast to the Formal Charge which divides each bonding pair equally without concern for which atom may be
more electronegative. When this division has been done for all bonds, the charge remaining on each atom is said to be its oxidation
number. If two like atoms are joined, each atom is assigned half the bonding electrons.
4.3.4 https://chem.libretexts.org/@go/page/119805
In each Lewis diagram, electrons have been color coded to indicate the atom from which they came originally. The boxes
enclose electrons assigned to a given atom by the rules for determining oxidation number.
a) Since the bond in Cl2 is purely covalent and the electrons are shared equally, one electron from the bond is assigned to each
Cl, giving the same number of valence electrons (7) as a neutral Cl atom. Thus neither atom has lost any electrons, and the
oxidation number is 0. This is indicated by writing a 0 above the symbol for chlorine in the formula
0
Cl 2 (4.3.2)
b) Since C is more electronegative than H, the pair of electrons in each C―H bond is assigned to C. Therefore each H has lost
the one valence electron it originally had, giving an oxidation number of +1. The C atom has gained four electrons, giving it a
negative charge and hence an oxidation number of – 4:
−4 +1
C H 4 (4.3.3)
c) In NaCl each Na atom has lost an electron to form an Na+ ion, and each Cl atom has gained an electron to form Cl–. The
oxidation numbers therefore correspond to the ionic charges:
+1 −1
Na Cl (4.3.4)
d) Since F is more electronegative than O, the bonding pairs are assigned to F in oxygen difluoride (OF2). The O is left with
four valence electrons, and each F has eight. The oxidation numbers are
+2 −1
O F2 (4.3.5)
e) In Hydrogen peroxide (H2O2) the O—H bond pairs are assigned to the more electronegative O’s, but the O―O bond is purely
covalent, and the electron pair is divided equally. This gives each O seven electrons, a gain of 1 over the neutral atom. The
oxidation numbers are
+1 −1
H2 O2 (4.3.6)
Oxidation numbers are mainly used by chemists to identify and handle a type of chemical reaction called a redox reaction, or an
oxidation-reduction reaction. This type of reaction can be recognized because it involves a change in oxidation number of at least
one element. We explored this last term. Oxidation numbers can sometimes also be useful in writing Lewis structures, particularly
for oxyanions. In the sulfite ion, SO32– for example, the oxidation number of sulfur is +4, suggesting that only four sulfur electrons
are involved in the bonding. Since sulfur has six valence electrons, we conclude that two electrons are not involved in the bonding,
i.e., that there is a lone pair. With this clue, a plausible Lewis structure is much easier to draw:
4.3.5 https://chem.libretexts.org/@go/page/119805
Oxidation Numbers and Lewis Structures
Summary
In a Lewis structure, formal charges can be assigned to each atom by treating each bond as if one-half of the electrons are assigned
to each atom. These hypothetical formal charges are a guide to determining the most appropriate Lewis structure. A structure in
which the formal charges are as close to zero as possible is preferred. Oxidation numbers can further be used to help elucidate the
correct Lewis structure.
Key Equations
1
formal charge = # valence shell electrons (free atom) − # one pair electrons − # bonding electrons
2
Glossary
formal charge
charge that would result on an atom by taking the number of valence electrons on the neutral atom and subtracting the
nonbonding electrons and the number of bonds (one-half of the bonding electrons)
molecular structure
arrangement of atoms in a molecule or ion
resonance
situation in which one Lewis structure is insufficient to describe the bonding in a molecule and the average of multiple
structures is observed
resonance forms
two or more Lewis structures that have the same arrangement of atoms but different arrangements of electrons
resonance hybrid
average of the resonance forms shown by the individual Lewis structures
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Ed Vitz (Kutztown University), John W. Moore (UW-Madison), Justin Shorb (Hope College), Xavier Prat-Resina (University of
Minnesota Rochester), Tim Wendorff, and Adam Hahn.
Adelaide Clark, Oregon Institute of Technology
4.3.6 https://chem.libretexts.org/@go/page/119805
Wayne Breslyn (video 1)
Clayton Spencer (video 2)
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
4.3: Formal Charge and Oxidation State is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
4.3.7 https://chem.libretexts.org/@go/page/119805
4.3: Formal Charge and Oxidation State (Problems)
PROBLEM 4.3.1
Determine the formal charge and oxidation state of each element in the following:
a. HCl
b. CF4
c. PCl3
Answer a
FC: H: 0, Cl: 0
OX: H: +1, Cl: -1
Answer b
FC: C: 0, F: 0
OX: C: +4, F: -1
Answer c
FC: P: 0, Cl: 0
OX: P: +3, Cl: -1
PROBLEM 4.3.2
Determine the formal charge and oxidation state of each element in the following:
a. H3O+
b. SO 2−
4
c. NH3
d. O2−
e. H2O2
Answer a
FC: H: 0, O: +1
OX: H: +1, O: -2
Answer b
FC: S: +2, O: -1
OX: S: +6, O: -2
Answer c
FC: N: 0, H: 0
OX: N: -3, H: +1
Answer d
FC: -1
OX: -1
Answer e
FC: O: 0, H: 0
OX: O: -1, H: +1
4.3.1 https://chem.libretexts.org/@go/page/119806
Click here to see a video of the solution
Problem 4.4.2
PROBLEM 4.3.3
Calculate the formal charge and oxidation state of chlorine in the molecules Cl2 and CCl4.
Answer
FC: Cl in Cl2: 0; Cl in CCl4: 0
OX: Cl in Cl2: 0; Cl in CCl4: -1
PROBLEM 4.3.4
Calculate the formal charge and oxidation state of each element in the following compounds and ions:
a. F2CO
b. NO–
c. BF−
4
d. SnCl −
e. H2CCH2
f. PO3−
Answer a
FC: C: 0, F :0, O: 0
OX: C: +4, F: -1, O: -2
Answer b
FC: N: -1, O: 0
OX: N: +1, O: -2
Answer c
FC: B: -1, F: 0
OX: B: +3, F: -1
Answer d
FC: Sn: -1; Cl: 0
OX: Sn: +2, Cl: -1
Answer e
FC: C: 0, H: 0
4.3.2 https://chem.libretexts.org/@go/page/119806
OX: C: -2, H: +1
Answer f
FC: P: +1, O: -1
OX: P: +5, O: -2
PROBLEM 4.3.5
Based on formal charge considerations, which of the following would likely be the correct arrangement of atoms in nitrosyl
chloride: ClNO or ClON?
Answer
ClNO
Problem 4.4.5
PROBLEM 4.3.6
Based on formal charge considerations, which of the following would likely be the correct arrangement of atoms in
hypochlorous acid: HOCl or OClH?
Answer
HOCl
PROBLEM 4.3.7
Based on formal charge considerations, which of the following would likely be the correct arrangement of atoms in sulfur
dioxide: OSO or SOO?
Answer
OSO
4.3.3 https://chem.libretexts.org/@go/page/119806
Problem 4.4.7
PROBLEM 4.3.8
Draw the structure of hydroxylamine, H3NO, and assign formal charges; look up the structure. Is the actual structure consistent
with the formal charges?
Answer
The structure that gives zero formal charges is consistent with the actual structure:
A Lewis structure shows a nitrogen atom with one lone pair of electrons single bonded to two hydrogen atoms and an oxygen atom which has two lone pairs of electrons. The oxygen atom is single bonded
to a hydrogen atom.
PROBLEM 4.3.9
Which of the following structures would we expect for nitrous acid? Determine the formal charges:
Two Lewis structures are shown, with the word “or” in between. The left structure shows a nitrogen atom single bonded to an oxygen atom with three lone pairs of electrons. It is also single bonded to a hydrogen
atom and double bonded to an oxygen atom with two lone pairs of electrons. The right structure shows a hydrogen atom single bonded to an oxygen atom with two lone pairs of electrons. The oxygen atom is single
bonded to a nitrogen atom which is double bonded to an oxygen atom with two lone pairs of electrons.
Answer
The first structure is the best structure. the formal charges are closest to 0 (and also the second structure does not give a
complete octet on N)
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Feedback
Think one of the answers above is wrong? Let us know here.
4.3: Formal Charge and Oxidation State (Problems) is shared under a CC BY license and was authored, remixed, and/or curated by LibreTexts.
4.3.4 https://chem.libretexts.org/@go/page/119806
CHAPTER OVERVIEW
Unit 5: The Strength and Shape of Covalent Bonds
Unit Objectives
Describe the energetics of covalent bond formation and breakage
Predict the structures of small molecules using valence shell electron pair repulsion (VSEPR) theory
Describe the formation of covalent bonds in terms of atomic orbital overlap
Define and give examples of σ and π bonds
Explain the concept of atomic orbital hybridization
Determine the hybrid orbitals associated with various molecular geometries
Unit 5: The Strength and Shape of Covalent Bonds is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.
1
5.1: Covalent Bond Formation and Strength
Skills to Develop
Describe the formation of covalent bonds
Describe the energetics of covalent bond formation and breakage
5.1.1 https://chem.libretexts.org/@go/page/112685
Figure 5.1.1 : The potential energy of two separate hydrogen atoms (right) decreases as they approach each other, and the single
electrons on each atom are shared to form a covalent bond. The bond length is the internuclear distance at which the lowest
potential energy is achieved.
It is essential to remember that energy must be added to break chemical bonds (an endothermic process), whereas forming chemical
bonds releases energy (an exothermic process). In the case of H2, the covalent bond is very strong; a large amount of energy, 436
kJ, must be added to break the bonds in one mole of hydrogen molecules and cause the atoms to separate:
Conversely, the same amount of energy is released when one mole of H2 molecules forms from two moles of H atoms:
The total number of electrons around each individual atom consists of six nonbonding electrons and two shared (i.e., bonding)
electrons for eight total electrons, matching the number of valence electrons in the noble gas argon. Since the bonding atoms are
identical, Cl2 also features a pure covalent bond.
When the atoms linked by a covalent bond are different, the bonding electrons are shared, but no longer equally. Instead, the
bonding electrons are more attracted to one atom than the other, giving rise to a shift of electron density toward that atom. This
unequal distribution of electrons is known as a polar covalent bond, characterized by a partial positive charge on one atom and a
partial negative charge on the other. The atom that attracts the electrons more strongly acquires the partial negative charge and vice
versa. For example, the electrons in the H–Cl bond of a hydrogen chloride molecule spend more time near the chlorine atom than
near the hydrogen atom. Thus, in an HCl molecule, the chlorine atom carries a partial negative charge and the hydrogen atom has a
partial positive charge. Figure 5.1.2 shows the distribution of electrons in the H–Cl bond. Note that the shaded area around Cl is
much larger than it is around H. Compare this to Figure 5.1.1, which shows the even distribution of electrons in the H2 nonpolar
bond.
Figure 5.1.2 : (a) The distribution of electron density in the HCl molecule is uneven. The electron density is greater around the
chlorine nucleus. The small, black dots indicate the location of the hydrogen and chlorine nuclei in the molecule. (b) Symbols δ+
and δ– indicate the polarity of the H–Cl bond.
We sometimes designate the positive and negative atoms in a polar covalent bond using a lowercase Greek letter “delta,” δ, with a
plus sign or minus sign to indicate whether the atom has a partial positive charge (δ+) or a partial negative charge (δ–). This
symbolism is shown for the H–Cl molecule in Figure 5.1.2b.
A bond’s strength describes how strongly each atom is joined to another atom, and therefore how much energy is required to break
the bond between the two atoms. In this section, you will learn about the bond strength of covalent bonds. Later in this course, we
will compare that to the strength of ionic bonds, which is related to the lattice energy of a compound.
5.1.2 https://chem.libretexts.org/@go/page/112685
Bond Strength: Covalent Bonds
Stable molecules exist because covalent bonds hold the atoms together. We measure the strength of a covalent bond by the energy
required to break it, that is, the energy necessary to separate the bonded atoms. Separating any pair of bonded atoms requires
energy; the stronger a bond, the greater the energy required to break it. The energy required to break a specific covalent bond in
one mole of gaseous molecules is called the bond energy or the bond dissociation energy. The bond energy for a diatomic molecule,
D X–Y , is defined as the standard enthalpy change for the endothermic reaction:
X Y(g) ⟶ X(g) + Y(g) DX−Y = ΔH ° (5.1.4)
For example, the bond energy of the pure covalent H–H bond, Δ H –H , is 436 kJ per mole of H–H bonds broken:
H2(g) ⟶ 2 H(g) DH −H = ΔH ° = 436kJ (5.1.5)
The average C–H bond energy, D C –H , is 1660/4 = 415 kJ/mol because there are four moles of C–H bonds broken per mole of the
reaction. Although the four C–H bonds are equivalent in the original molecule, they do not each require the same energy to break;
once the first bond is broken (which requires 439 kJ/mol), the remaining bonds are easier to break. The 415 kJ/mol value is the
average, not the exact value required to break any one bond.
The strength of a bond between two atoms increases as the number of electron pairs in the bond increases. Generally, as the bond
strength increases, the bond length decreases. Thus, we find that triple bonds are stronger and shorter than double bonds between
the same two atoms; likewise, double bonds are stronger and shorter than single bonds between the same two atoms. Average bond
energies for some common bonds appear in Table 5.1.2, and a comparison of bond lengths and bond strengths for some common
bonds appears in Table 5.1.2. When one atom bonds to various atoms in a group, the bond strength typically decreases as we move
down the group. For example, C–F is 439 kJ/mol, C–Cl is 330 kJ/mol, and C–Br is 275 kJ/mol.
Table 5.1.1: Bond Energies (kJ/mol)
Bond Bond Energy Bond Bond Energy Bond Bond Energy
5.1.3 https://chem.libretexts.org/@go/page/112685
Bond Bond Energy Bond Bond Energy Bond Bond Energy
Table 5.1.2: Average Bond Lengths and Bond Energies for Some Common Bonds
Bond Bond Length (Å) Bond Energy (kJ/mol)
C = C 1.34 611
C ≡ C 1.20 837
C = N 1.38 615
C ≡ N 1.16 891
C = O 1.23 741
C ≡ O 1.13 1080
We can use bond energies to calculate approximate enthalpy changes for reactions where enthalpies of formation are not available.
Calculations of this type will also tell us whether a reaction is exothermic or endothermic.
An exothermic reaction (ΔH negative, heat produced) results when the bonds in the products are stronger than the bonds in
the reactants.
An endothermic reaction (ΔH positive, heat absorbed) results when the bonds in the products are weaker than those in the
reactants.
The enthalpy change, ΔH, for a chemical reaction is approximately equal to the sum of the energy required to break all bonds in the
reactants (energy “in”, positive sign) plus the energy released when all bonds are formed in the products (energy “out,” negative
sign). This can be expressed mathematically in the following way:
In this expression, the symbol Σ means “the sum of” and D represents the bond energy in kilojoules per mole, which is always a
positive number. The bond energy is obtained from a table and will depend on whether the particular bond is a single, double, or
triple bond. Thus, in calculating enthalpies in this manner, it is important that we consider the bonding in all reactants and products.
5.1.4 https://chem.libretexts.org/@go/page/112685
Because D values are typically averages for one type of bond in many different molecules, this calculation provides a rough
estimate, not an exact value, for the enthalpy of reaction.
Consider the following reaction:
H + Cl ⟶ 2 HCl (5.1.7)
2 (g) 2 (g) (g)
or
H– H + Cl– Cl ⟶ 2 H– Cl (5.1.8)
(g) (g) (g)
To form two moles of HCl, one mole of H–H bonds and one mole of Cl–Cl bonds must be broken. The energy required to break
these bonds is the sum of the bond energy of the H–H bond (436 kJ/mol) and the Cl–Cl bond (243 kJ/mol). During the reaction,
two moles of H–Cl bonds are formed (bond energy = 432 kJ/mol), releasing 2 × 432 kJ; or 864 kJ. Because the bonds in the
products are stronger than those in the reactants, the reaction releases more energy than it consumes:
This excess energy is released as heat, so the reaction is exothermic. Table T2 gives a value for the standard molar enthalpy of
formation of HCl(g), ΔH , of –92.307 kJ/mol. Twice that value is –184.6 kJ, which agrees well with the answer obtained earlier
∘
f
Solution
First, we need to write the Lewis structures of the reactants and the products:
From this, we see that ΔH for this reaction involves the energy required to break a C–O triple bond and two H–H single bonds,
as well as the energy produced by the formation of three C–H single bonds, a C–O single bond, and an O–H single bond. We
can express this as follows (via Equation 5.1.6):
= −107 kJ
We can compare this value to the value calculated based on ΔH data from Appendix G:
f
∘
∘ ∘ ∘
ΔH = [ΔH CH OH(g)] − [ΔH CO(g) + 2 × ΔH H ]
f 3 f f 2
= [−201.0] − [−110.52 + 2 × 0]
= −90.5 kJ
Note that there is a fairly significant gap between the values calculated using the two different methods. This occurs because D
values are the average of different bond strengths; therefore, they often give only rough agreement with other data.
5.1.5 https://chem.libretexts.org/@go/page/112685
Exercise 5.1.1
Ethyl alcohol, CH3CH2OH, was one of the first organic chemicals deliberately synthesized by humans. It has many uses in
industry, and it is the alcohol contained in alcoholic beverages. It can be obtained by the fermentation of sugar or synthesized by
the hydration of ethylene in the following reaction:
Using the bond energies in Table 5.1.2, calculate an approximate enthalpy change, ΔH, for this reaction.
Answer
–35 kJ
Summary
Covalent bonds form when electrons are shared between atoms and are attracted by the nuclei of both atoms. In pure covalent
bonds, the electrons are shared equally. In polar covalent bonds, the electrons are shared unequally, as one atom exerts a stronger
force of attraction on the electrons than the other.
The strength of a covalent bond is measured by its bond dissociation energy, that is, the amount of energy required to break that
particular bond in a mole of molecules. Multiple bonds are stronger than single bonds between the same atoms. The enthalpy of a
reaction can be estimated based on the energy input required to break bonds and the energy released when new bonds are formed.
Key Equations
Bond energy for a diatomic molecule: XY(g) ⟶ X(g) + Y(g) D
X–Y
= ΔH °
Footnotes
1. This question is taken from the Chemistry Advanced Placement Examination and is used with the permission of the Educational
Testing Service.
5.1.6 https://chem.libretexts.org/@go/page/112685
Glossary
bond energy
(also, bond dissociation energy) energy required to break a covalent bond in a gaseous substance
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Fuse School, Open Educational Resource free of charge, under a Creative Commons License: Attribution-NonCommercial CC
BY-NC (View License Deed: https://creativecommons.org/licenses/by-nc/4.0/)
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
5.1: Covalent Bond Formation and Strength is shared under a CC BY license and was authored, remixed, and/or curated by LibreTexts.
5.1.7 https://chem.libretexts.org/@go/page/112685
5.1: Covalent Bond Formation and Strength (Problems)
PROBLEM 5.1.1
Which bond in each of the following pairs of bonds is the strongest?
a. C–C or C = C
b. C–N or C ≡ N
c. C ≡ O or C = O
d. H–F or H–Cl
e. C–H or O–H
f. C–N or C–O
Answer a
C =C
Answer b
C ≡N
Answer c
C ≡O
Answer d
H–F
Answer e
O–H
Answer f
C–O
PROBLEM 5.1.2
When a molecule can form two different structures, the structure with the stronger bonds is usually the more stable form. Use
bond energies to predict the correct structure of the hydroxylamine molecule:
Two Lewis structures are shows with the word “or” written in between them. The left structure shows a nitrogen atom with one lone pair of electrons single bonded to two hydrogen atoms. It is also bonded to an
oxygen atom with two lone pairs of electrons. The oxygen atom is bonded to a hydrogen atom. The right structure shows a nitrogen atom single bonded to three hydrogen atoms and an oxygen atom with three lone
pairs of electrons.
Answer
The greater bond energy is in the figure on the left. It is the more stable form.
PROBLEM 5.1.3
Use principles of atomic structure to answer each of the following:1
a. The radius of the Ca atom is 197 pm; the radius of the Ca2+ ion is 99 pm. Account for the difference.
b. Given these ionization values, explain the difference between Ca and K with regard to their first and second ionization
energies.
K 419 3050
Ca 590 1140
5.1.1 https://chem.libretexts.org/@go/page/112688
c. The first ionization energy of Mg is 738 kJ/mol and that of Al is 578 kJ/mol. Account for this difference.
Answer a
When two electrons are removed from the valence shell, the Ca radius loses the outermost energy level and reverts to the
lower n = 3 level, which is much smaller in radius.
Answer b
Removal of the 4s electron in Ca requires more energy than removal of the 4s electron in K because of the stronger attraction
of the nucleus and the extra energy required to break the pairing of the electrons. The second ionization energy for K
requires that an electron be removed from a lower energy level, where the attraction is much stronger from the nucleus for
the electron. In addition, energy is required to unpair two electrons in a full orbital. For Ca, the second ionization potential
requires removing only a lone electron in the exposed outer energy level.
Answer c
In Al, the removed electron is relatively unprotected and unpaired in a p orbital. The higher energy for Mg mainly reflects
the unpairing of the 2s electron
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Feedback
Think one of the answers above is wrong? Let us know here.
5.1: Covalent Bond Formation and Strength (Problems) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated
by LibreTexts.
5.1.2 https://chem.libretexts.org/@go/page/112688
5.2: Molecular Shape
Skills to Develop
Predict the structures of small molecules using valence shell electron pair repulsion (VSEPR) theory
Thus far, we have used two-dimensional Lewis structures to represent molecules. However, molecular structure is actually three-
dimensional, and it is important to be able to describe molecular bonds in terms of their distances, angles, and relative
arrangements in space (Figure 5.2.1). A bond angle is the angle between any two bonds that include a common atom, usually
measured in degrees. A bond distance (or bond length) is the distance between the nuclei of two bonded atoms along the straight
line joining the nuclei. Bond distances are measured in Ångstroms (1 Å = 10–10 m) or picometers (1 pm = 10–12 m, 100 pm = 1 Å).
Figure 5.2.1 : Bond distances (lengths) and angles are shown for the formaldehyde molecule, H2CO.
VSEPR Theory
Valence shell electron-pair repulsion theory (VSEPR theory) enables us to predict the molecular structure, including approximate
bond angles around a central atom, of a molecule from an examination of the number of bonds and lone electron pairs in its Lewis
structure. The VSEPR model assumes that electron pairs in the valence shell of a central atom will adopt an arrangement that
minimizes repulsions between these electron pairs by maximizing the distance between them. The electrons in the valence shell of
a central atom form either bonding pairs of electrons, located primarily between bonded atoms, or lone pairs. The electrostatic
repulsion of these electrons is reduced when the various regions of high electron density assume positions as far from each other as
possible.
VSEPR theory predicts the arrangement of electron pairs around each central atom and, usually, the correct arrangement of atoms
in a molecule. We should understand, however, that the theory only considers electron-pair repulsions. Other interactions, such as
nuclear-nuclear repulsions and nuclear-electron attractions, are also involved in the final arrangement that atoms adopt in a
particular molecular structure.
As a simple example of VSEPR theory, let us predict the structure of a gaseous BeF2 molecule. The Lewis structure of BeF2
(Figure 5.2.2) shows only two electron pairs around the central beryllium atom. With two bonds and no lone pairs of electrons on
the central atom, the bonds are as far apart as possible, and the electrostatic repulsion between these regions of high electron
density is reduced to a minimum when they are on opposite sides of the central atom. The bond angle is 180° (Figure 5.2.2).
Figure 5.2.2 : The BeF2 molecule adopts a linear structure in which the two bonds are as far apart as possible, on opposite sides
of the Be atom.
Figure 5.2.3 illustrates this and other electron-pair geometries that minimize the repulsions among regions of high electron density
(bonds and/or lone pairs). Two regions of electron density around a central atom in a molecule form a linear geometry; three
regions form a trigonal planar geometry; four regions form a tetrahedral geometry; five regions form a trigonal bipyramidal
geometry; and six regions form an octahedral geometry.
5.2.1 https://chem.libretexts.org/@go/page/112686
Figure 5.2.3 : The basic electron-pair geometries predicted by VSEPR theory maximize the space around any region of electron
density (bonds or lone pairs).
Figure 5.2.4 : The molecular structure of the methane molecule, CH4, is shown with a tetrahedral arrangement of the hydrogen
atoms. VSEPR structures like this one are often drawn using the wedge and dash notation, in which solid lines represent bonds in
the plane of the page, solid wedges represent bonds coming up out of the plane, and dashed lines represent bonds going down into
the plane.
For example, the methane molecule, CH4, which is the major component of natural gas, has four bonding pairs of electrons around
the central carbon atom; the electron-pair geometry is tetrahedral, as is the molecular structure (Figure 5.2.4). On the other hand,
the ammonia molecule, NH3, also has four electron pairs associated with the nitrogen atom, and thus has a tetrahedral electron-pair
5.2.2 https://chem.libretexts.org/@go/page/112686
geometry. One of these regions, however, is a lone pair, which is not included in the molecular structure, and this lone pair
influences the shape of the molecule (Figure 5.2.5).
Figure 5.2.5 : (a) The electron-pair geometry for the ammonia molecule is tetrahedral with one lone pair and three single bonds.
(b) The trigonal pyramidal molecular structure is determined from the electron-pair geometry. (c) The actual bond angles deviate
slightly from the idealized angles because the lone pair takes up a larger region of space than do the single bonds, causing the
HNH angle to be slightly smaller than 109.5°.
Small distortions from the ideal angles in Figure 5.2.5 can result from differences in repulsion between various regions of electron
density. VSEPR theory predicts these distortions by establishing an order of repulsions and an order of the amount of space
occupied by different kinds of electron pairs. The order of electron-pair repulsions from greatest to least repulsion is:
lone pair-lone pair > lone pair-bonding pair > bonding pair-bonding pair
This order of repulsions determines the amount of space occupied by different regions of electrons. A lone pair of electrons
occupies a larger region of space than the electrons in a triple bond; in turn, electrons in a triple bond occupy more space than those
in a double bond, and so on. The order of sizes from largest to smallest is:
lone pair > triple bond > double bond > single bond
Consider formaldehyde, H2CO, which is used as a preservative for biological and anatomical specimens. This molecule has regions
of high electron density that consist of two single bonds and one double bond. The basic geometry is trigonal planar with 120°
bond angles, but we see that the double bond causes slightly larger angles (121°), and the angle between the single bonds is slightly
smaller (118°).
In the ammonia molecule, the three hydrogen atoms attached to the central nitrogen are not arranged in a flat, trigonal planar
molecular structure, but rather in a three-dimensional trigonal pyramid (Figure 5.2.6) with the nitrogen atom at the apex and the
three hydrogen atoms forming the base. The ideal bond angles in a trigonal pyramid are based on the tetrahedral electron pair
geometry. Again, there are slight deviations from the ideal because lone pairs occupy larger regions of space than do bonding
electrons. The H–N–H bond angles in NH3 are slightly smaller than the 109.5° angle in a regular tetrahedron (Figure 5.2.6)
because the lone pair-bonding pair repulsion is greater than the bonding pair-bonding pair repulsion. The ideal molecular structures
are predicted based on the electron-pair geometries for various combinations of lone pairs and bonding pairs.
5.2.3 https://chem.libretexts.org/@go/page/112686
Figure 5.2.6 : The molecular structures are identical to the electron-pair geometries when there are no lone pairs present (first
column). For a particular number of electron pairs (row), the molecular structures for one or more lone pairs are determined
based on modifications of the corresponding electron-pair geometry.
According to VSEPR theory, the terminal atom locations (Xs in Figure 5.2.7) are equivalent within the linear, trigonal planar, and
tetrahedral electron-pair geometries (the first three rows of the table). It does not matter which X is replaced with a lone pair
because the molecules can be rotated to convert positions. For trigonal bipyramidal electron-pair geometries, however, there are
two distinct X positions (Figure 5.2.7a): an axial position (if we hold a model of a trigonal bipyramid by the two axial positions,
we have an axis around which we can rotate the model) and an equatorial position (three positions form an equator around the
middle of the molecule). The axial position is surrounded by bond angles of 90°, whereas the equatorial position has more space
available because of the 120° bond angles. In a trigonal bipyramidal electron-pair geometry, lone pairs always occupy equatorial
positions because these more spacious positions can more easily accommodate the larger lone pairs.
Theoretically, we can come up with three possible arrangements for the three bonds and two lone pairs for the ClF3 molecule
(Figure 5.2.7). The stable structure is the one that puts the lone pairs in equatorial locations, giving a T-shaped molecular structure.
Figure 5.2.7 : (a) In a trigonal bipyramid, the two axial positions are located directly across from one another, whereas the three
equatorial positions are located in a triangular arrangement. (b–d) The two lone pairs (red lines) in ClF3 have several possible
5.2.4 https://chem.libretexts.org/@go/page/112686
arrangements, but the T-shaped molecular structure (b) is the one actually observed, consistent with the larger lone pairs both
occupying equatorial positions.
When a central atom has two lone electron pairs and four bonding regions, we have an octahedral electron-pair geometry. The two
lone pairs are on opposite sides of the octahedron (180° apart), giving a square planar molecular structure that minimizes lone pair-
lone pair repulsions.
Predicting Electron Pair Geometry and Molecular Structure
The following procedure uses VSEPR theory to determine the electron pair geometries and the molecular structures:
1. Write the Lewis structure of the molecule or polyatomic ion.
2. Count the number of regions of electron density (lone pairs and bonds) around the central atom. A single, double, or triple
bond counts as one region of electron density.
3. Identify the electron-pair geometry based on the number of regions of electron density: linear, trigonal planar, tetrahedral,
trigonal bipyramidal, or octahedral (Figure 5.2.7, first column).
4. Use the number of lone pairs to determine the molecular structure (Figure 5.2.7 ). If more than one arrangement of lone pairs
and chemical bonds is possible, choose the one that will minimize repulsions, remembering that lone pairs occupy more
space than multiple bonds, which occupy more space than single bonds. In trigonal bipyramidal arrangements, repulsion is
minimized when every lone pair is in an equatorial position. In an octahedral arrangement with two lone pairs, repulsion is
minimized when the lone pairs are on opposite sides of the central atom.
The following examples illustrate the use of VSEPR theory to predict the molecular structure of molecules or ions that have no
lone pairs of electrons. In this case, the molecular structure is identical to the electron pair geometry.
This shows us two regions of high electron density around the carbon atom—each double bond counts as one region, and there
are no lone pairs on the carbon atom. Using VSEPR theory, we predict that the two regions of electron density arrange
themselves on opposite sides of the central atom with a bond angle of 180°. The electron-pair geometry and molecular structure
are identical, and CO2 molecules are linear.
(b) We write the Lewis structure of BCl3 as:
Thus we see that BCl3 contains three bonds, and there are no lone pairs of electrons on boron. The arrangement of three regions
of high electron density gives a trigonal planar electron-pair geometry. The B–Cl bonds lie in a plane with 120° angles between
them. BCl3 also has a trigonal planar molecular structure.
The electron-pair geometry and molecular structure of BCl3 are both trigonal planar. Note that the VSEPR geometry indicates
the correct bond angles (120°), unlike the Lewis structure shown above.
5.2.5 https://chem.libretexts.org/@go/page/112686
Exercise 5.2.1
Carbonate, CO , is a common polyatomic ion found in various materials from eggshells to antacids. What are the electron-pair
2−
Answer
The electron-pair geometry is trigonal planar and the molecular structure is trigonal planar. Due to resonance, all three C–O
bonds are identical. Whether they are single, double, or an average of the two, each bond counts as one region of electron
density.
Solution
We write the Lewis structure of NH as:
+
4
We can see that NH contains four bonds from the nitrogen atom to hydrogen atoms and no lone pairs. We expect the four
+
regions of high electron density to arrange themselves so that they point to the corners of a tetrahedron with the central nitrogen
atom in the middle (Figure 5.2.7). Therefore, the electron pair geometry of NH is tetrahedral, and the molecular structure is
+
Figure 5.2.8 : The ammonium ion displays a tetrahedral electron-pair geometry as well as a tetrahedral molecular structure.
Exercise 5.2.2
Identify a molecule with trigonal bipyramidal molecular structure.
Answer
Any molecule with five electron pairs around the central atoms including no lone pairs will be trigonal bipyramidal. PF is a 5
common example
The next several examples illustrate the effect of lone pairs of electrons on molecular structure.
5.2.6 https://chem.libretexts.org/@go/page/112686
We predict that these four regions are arranged in a tetrahedral fashion (Figure 5.2.6), as indicated in Figure 5.2.9. Thus, the
electron-pair geometry is tetrahedral and the molecular structure is bent with an angle slightly less than 109.5°. In fact, the bond
angle is 104.5°.
Figure 5.2.9 : (a) H2O has four regions of electron density around the central atom, so it has a tetrahedral electron-pair
geometry. (b) Two of the electron regions are lone pairs, so the molecular structure is bent.
Exercise 5.2.3
The hydronium ion, H3O+, forms when acids are dissolved in water. Predict the electron-pair geometry and molecular structure
of this cation.
Answer
electron pair geometry: tetrahedral; molecular structure: trigonal pyramidal
We expect these five regions to adopt a trigonal bipyramidal electron-pair geometry. To minimize lone pair repulsions, the lone
pair occupies one of the equatorial positions. The molecular structure (Figure 5.2.6) is that of a seesaw (Figure 5.2.10).
5.2.7 https://chem.libretexts.org/@go/page/112686
Figure 5.2.10 : (a) SF4 has a trigonal bipyramidal arrangement of the five regions of electron density. (b) One of the regions is
a lone pair, which results in a seesaw-shaped molecular structure.
Exercise 5.2.4
Predict the electron pair geometry and molecular structure for molecules of XeF2.
Answer
The electron-pair geometry is trigonal bipyramidal. The molecular structure is linear.
These six regions adopt an octahedral arrangement (Figure 5.2.6), which is the electron-pair geometry. To minimize repulsions,
the lone pairs should be on opposite sides of the central atom (Figure 5.2.11). The five atoms are all in the same plane and have
a square planar molecular structure.
Figure 5.2.11 : (a) XeF4 adopts an octahedral arrangement with two lone pairs (red lines) and four bonds in the electron-pair
geometry. (b) The molecular structure is square planar with the lone pairs directly across from one another.
Exercise 5.2.4
5.2.8 https://chem.libretexts.org/@go/page/112686
In a certain molecule, the central atom has three lone pairs and two bonds. What will the electron pair geometry and molecular
structure be?
Answer
electron pair geometry: trigonal bipyramidal; molecular structure: linear
Solution
A Lewis structure depicts a nitrogen atom with one lone pair of electrons that is single bonded to two hydrogen atoms and a carbon atom. The atoms described are drawn with bonds that indicate a three-
dimensional, tetrahedral shape around the nitrogen atom. The carbon is, in turn, single bonded to two hydrogen atoms and another carbon atom, and again, a tetrahedral, three dimensional configuration is indicated
by the types of bonds. This second carbon atom is double bonded to an oxygen atom and single bonded to an oxygen that has two lone pairs of electrons and a single bond to a hydrogen atom.
5.2.9 https://chem.libretexts.org/@go/page/112686
oxygen (OH)—two bonds, two lone pairs; bent (109°)
Exercise 5.2.5
Another amino acid is alanine, which has the Lewis structure shown here. Predict the electron-pair geometry and local structure
of the nitrogen atom, the three carbon atoms, and the oxygen atom with hydrogen attached:
Answer
electron-pair geometries: nitrogen––tetrahedral; carbon (CH)—tetrahedral; carbon (CH3)—tetrahedral; carbon (CO2)—
trigonal planar; oxygen (OH)—tetrahedral; local structures: nitrogen—trigonal pyramidal; carbon (CH)—tetrahedral; carbon
(CH3)—tetrahedral; carbon (CO2)—trigonal planar; oxygen (OH)—bent (109°)
5.2.10 https://chem.libretexts.org/@go/page/112686
Molecule Shapes
O
A
H
X X
Real Molecul
Model
Using this molecular shape simulator allows us to control whether bond angles and/or lone pairs are displayed by checking or
unchecking the boxes under “Options” on the right. We can also use the “Name” checkboxes at bottom-left to display or hide
the electron pair geometry (called “electron geometry” in the simulator) and/or molecular structure (called “molecular shape” in
the simulator).
Build the molecule HCN in the simulator based on the following Lewis structure:
H– C ≡ N
Click on each bond type or lone pair at right to add that group to the central atom. Once you have the complete molecule, rotate
it to examine the predicted molecular structure. What molecular structure is this?
Solution
The molecular structure is linear.
Exercise 5.2.6
Build a more complex molecule in the simulator. Identify the electron-group geometry, molecular structure, and bond angles.
Then try to find a chemical formula that would match the structure you have drawn.
5.2.11 https://chem.libretexts.org/@go/page/112686
Answer
Answers will vary. For example, an atom with four single bonds, a double bond, and a lone pair has an octahedral electron-
group geometry and a square pyramidal molecular structure. XeOF4 is a molecule that adopts this structure.
Summary
Glossary
axial position
location in a trigonal bipyramidal geometry in which there is another atom at a 180° angle and the equatorial positions are at a
90° angle
bond angle
angle between any two covalent bonds that share a common atom
bond distance
(also, bond length) distance between the nuclei of two bonded atoms
dipole moment
property of a molecule that describes the separation of charge determined by the sum of the individual bond moments based on
the molecular structure
electron-pair geometry
arrangement around a central atom of all regions of electron density (bonds, lone pairs, or unpaired electrons)
equatorial position
one of the three positions in a trigonal bipyramidal geometry with 120° angles between them; the axial positions are located at a
90° angle
linear
5.2.12 https://chem.libretexts.org/@go/page/112686
shape in which two outside groups are placed on opposite sides of a central atom
molecular structure
structure that includes only the placement of the atoms in the molecule
octahedral
shape in which six outside groups are placed around a central atom such that a three-dimensional shape is generated with four
groups forming a square and the other two forming the apex of two pyramids, one above and one below the square plane
polar molecule
(also, dipole) molecule with an overall dipole moment
tetrahedral
shape in which four outside groups are placed around a central atom such that a three-dimensional shape is generated with four
corners and 109.5° angles between each pair and the central atom
trigonal bipyramidal
shape in which five outside groups are placed around a central atom such that three form a flat triangle with 120° angles
between each pair and the central atom, and the other two form the apex of two pyramids, one above and one below the
triangular plane
trigonal planar
shape in which three outside groups are placed in a flat triangle around a central atom with 120° angles between each pair and
the central atom
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Fuse School, Open Educational Resource free of charge, under a Creative Commons License: Attribution-NonCommercial CC
BY-NC (View License Deed: https://creativecommons.org/licenses/by-nc/4.0/)
TED-Ed’s commitment to creating lessons worth sharing is an extension of TED’s mission of spreading great ideas. Within
TED-Ed’s growing library of TED-Ed animations, you will find carefully curated educational videos, many of which represent
collaborations between talented educators and animators nominated through the TED-Ed website.
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
5.2: Molecular Shape is shared under a CC BY license and was authored, remixed, and/or curated by LibreTexts.
5.2.13 https://chem.libretexts.org/@go/page/112686
5.2: Molecular Shape (Problems)
PROBLEM 5.2.1
Explain why the HOH molecule is bent, whereas the HBeH molecule is linear.
Answer
The placement of the two sets of unpaired electrons in water forces the bonds to assume a tetrahedral arrangement, and the
resulting HOH molecule is bent. The HBeH molecule (in which Be has only two electrons to bond with the two electrons
from the hydrogens) must have the electron pairs as far from one another as possible and is therefore linear.
PROBLEM 5.2.2
Explain the difference between electron-pair geometry and molecular structure.
Answer
Space must be provided for each pair of electrons whether they are in a bond or are present as lone pairs. Electron-pair
geometry considers the placement of all electrons. Molecular structure considers only the bonding-pair geometry.
PROBLEM 5.2.3
Why is the H–N–H angle in NH3 smaller than the H–C–H bond angle in CH4? Why is the H–N–H angle in NH
+
4
identical to
the H–C–H bond angle in CH4?
Answer
NH3 has a lone pair of electrons, which forces the bonds closer together, lessening the bond angle compared to species with
4 bonds and no lone pairs.
PROBLEM 5.2.4
Predict the electron pair geometry and the molecular structure of each of the following molecules or ions:
a. BeH2 (hint: Be does not have a complete octet)
b. CH (hint: C does ot have a complete octet)
+
3
Answer a
Both the electron geometry and the molecular structure are linear.
Answer b
Both the electron geometry and the molecular structure are trigonal planar.
Click here to see a video of the solution
Problem 5.2.4
5.2.1 https://chem.libretexts.org/@go/page/112689
PROBLEM 5.2.5
Identify the electron pair geometry and the molecular structure of each of the following molecules or ions:
a. CF4
b. BF3
c. BeCl2
Answer a
Both the electron geometry and the molecular structure are tetrahedral.
Answer b
Both the electron geometry and the molecular structure are trigonal planar.
Answer c
Both the electron geometry and the molecular structure are linear.
PROBLEM 5.2.6
What are the electron-pair geometry and the molecular structure of each of the following molecules or ions?
a. ClO −
b. PCl3
c. PH −
Answer a
electron-pair geometry: tetrahedral
molecular structure: bent
Answer b
electron-pair geometry: tetrahedral
molecular structure: trigonal pyramidal
Answer c
electron-pair geometry: tetrahedral
molecular structure: bent (109°)
Click here to see a video of the solution
Problem 5.2.6
PROBLEM 5.2.7
5.2.2 https://chem.libretexts.org/@go/page/112689
Identify the electron pair geometry and the molecular structure of each of the following molecules:
a. ClNO (N is the central atom)
b. CS2
c. Cl2CO (C is the central atom)
d. Cl2SO (S is the central atom)
e. SO2F2 (S is the central atom)
f. (g) ClOF (Cl is the central atom)
+
2
Answer a
electron-pair geometry: trigonal planar, molecular structure: bent (120°)
Answer b
electron-pair geometry: linear, molecular structure: linear
Answer c
electron-pair geometry: trigonal planar, molecular structure: trigonal planar
Answer d
electron-pair geometry: tetrahedral, molecular structure: trigonal pyramidal
Answer e
electron-pair geometry: tetrahedral, molecular structure: tetrahedral
Answer f
electron-pair geometry: tetrahedral, molecular structure: trigonal pyramidal
PROBLEM 5.2.8
Draw the Lewis structures and predict the shape of each compound or ion:
a. CO2
b. NO −
c. SO3
d. SO2−
Answer a
linear
Answer b
Answer c
5.2.3 https://chem.libretexts.org/@go/page/112689
trigonal planar
Answer d
Problem 5.2.8
PROBLEM 5.2.9
A molecule with the formula AB2, in which A and B represent different atoms, could have one of three different shapes. Sketch
and name the three different shapes that this molecule might have. Give an example of a molecule or ion for each shape.
Answer
Three Lewis diagrams are shown. The first diagram shows the letter A single bonded to the left and right to the letter B. An example, “C O subscript 2,” and the term, “linear,” are written beside this
diagram. The second diagram shows the letter A with two lone pairs of electrons, single bonded to the left and lower right to the letter B. An example, “H subscript 2 O,” and the term, “bent with an
approximately 109 degree angle,” are written beside this diagram. The third diagram shows the letter A with one lone electron pair, single bonded to the left and lower right to the letter B. An example, “S O
subscript 2,” and the term, “bent with an approximately 120 degree angle,” are written beside this diagram.
PROBLEM 5.2.10
A molecule with the formula AB3, in which A and B represent different atoms, could have one of two different shapes. Sketch
and name the three different shapes that this molecule might have. Give an example of a molecule or ion that has each shape.
Answer
5.2.4 https://chem.libretexts.org/@go/page/112689
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Feedback
Think one of the answers above is wrong? Let us know here.
5.2: Molecular Shape (Problems) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
5.2.5 https://chem.libretexts.org/@go/page/112689
5.3: Valence Bond Theory and Hybrid Orbitals
Skills to Develop
Describe the formation of covalent bonds in terms of atomic orbital overlap
Define and give examples of σ and π bonds
Explain the concept of atomic orbital hybridization
Determine the hybrid orbitals associated with various molecular geometries
5.3.1 https://chem.libretexts.org/@go/page/126655
which varies depending on the atoms involved, the energy reaches its lowest (most stable) value. This optimum distance between
the two bonded nuclei is the bond distance between the two atoms. The bond is stable because at this point, the attractive and
repulsive forces combine to create the lowest possible energy configuration. If the distance between the nuclei were to decrease
further, the repulsions between nuclei and the repulsions as electrons are confined in closer proximity to each other would become
stronger than the attractive forces. The energy of the system would then rise (making the system destabilized), as shown at the far
left of Figure 5.3.1.
Figure 5.3.1 : The interaction of two hydrogen atoms changes as a function of distance. The energy of the system changes as the
atoms interact. The lowest (most stable) energy occurs at a distance of 74 pm, which is the bond length observed for the H2
molecule.
The bond energy is the difference between the energy minimum (which occurs at the bond distance) and the energy of the two
separated atoms. This is the quantity of energy released when the bond is formed. Conversely, the same amount of energy is
required to break the bond. For the H molecule shown in Figure 5.3.1, at the bond distance of 74 pm the system is
2
J lower in energy than the two separated hydrogen atoms. This may seem like a small number. However, we know
−19
7.24 × 10
from our earlier description of thermochemistry that bond energies are often discussed on a per-mole basis. For example, it requires
J to break one H–H bond, but it takes 4.36 × 10 J to break 1 mole of H–H bonds. A comparison of some bond
−19 5
7.24 × 10
lengths and energies is shown in Table 5.3.1. We can find many of these bonds in a variety of molecules, and this table provides
average values. For example, breaking the first C–H bond in CH4 requires 439.3 kJ/mol, while breaking the first C–H bond in
H– CH C H (a common paint thinner) requires 375.5 kJ/mol.
2 6 5
5.3.2 https://chem.libretexts.org/@go/page/126655
Bond Length (pm) Energy (kJ/mol) Bond Length (pm) Energy (kJ/mol)
In addition to the distance between two orbitals, the orientation of orbitals also affects their overlap (other than for two s orbitals,
which are spherically symmetric). Greater overlap is possible when orbitals are oriented such that they overlap on a direct line
between the two nuclei. Figure 5.3.2 illustrates this for two p orbitals from different atoms; the overlap is greater when the orbitals
overlap end to end rather than at an angle.
Figure 5.3.2 : (a) The overlap of two p orbitals is greatest when the orbitals are directed end to end. (b) Any other arrangement
results in less overlap. The dots indicate the locations of the nuclei.
The overlap of two s orbitals (as in H2), the overlap of an s orbital and a p orbital (as in HCl), and the end-to-end overlap of two p
orbitals (as in Cl2) all produce sigma bonds (σ bonds), as illustrated in Figure 5.3.3. A σ bond is a covalent bond in which the
electron density is concentrated in the region along the internuclear axis; that is, a line between the nuclei would pass through the
center of the overlap region. Single bonds in Lewis structures are described as σ bonds in valence bond theory.
Figure 5.3.3 : Sigma (σ) bonds form from the overlap of the following: (a) two s orbitals, (b) an s orbital and a p orbital, and (c)
two p orbitals. The dots indicate the locations of the nuclei.
A pi bond (π bond) is a type of covalent bond that results from the side-by-side overlap of two p orbitals, as illustrated in Figure
5.3.4. In a π bond, the regions of orbital overlap lie on opposite sides of the internuclear axis. Along the axis itself, there is a node,
Figure 5.3.4 : Pi (π) bonds form from the side-by-side overlap of two p orbitals. The dots indicate the location of the nuclei.
While all single bonds are σ bonds, multiple bonds consist of both σ and π bonds. As the Lewis structures suggest, O2 contains a
double bond, and N2 contains a triple bond. The double bond consists of one σ bond and one π bond, and the triple bond consists of
one σ bond and two π bonds. Between any two atoms, the first bond formed will always be a σ bond, but there can only be one σ
bond in any one location. In any multiple bond, there will be one σ bond, and the remaining one or two bonds will be π bonds.
These bonds are described in more detail later in this chapter.
As seen in Table 5.3.1, an average carbon-carbon single bond is 347 kJ/mol, while in a carbon-carbon double bond, the π bond
increases the bond strength by 267 kJ/mol. Adding an additional π bond causes a further increase of 225 kJ/mol. We can see a
5.3.3 https://chem.libretexts.org/@go/page/126655
similar pattern when we compare other σ and π bonds. Thus, each individual π bond is generally weaker than a corresponding σ
bond between the same two atoms. In a σ bond, there is a greater degree of orbital overlap than in a π bond.
Butadiene
Solution
There are six σ C–H bonds and one σ C–C bond, for a total of seven from the single bonds. There are two double bonds that
each have a π bond in addition to the σ bond. This gives a total nine σ and two π bonds overall.
Exercise 5.3.1
Identify each illustration as depicting a σ or π bond:
a. side-by-side overlap of a 4p and a 2p orbital
b. end-to-end overlap of a 4p and 4p orbital
c. end-to-end overlap of a 4p and a 2p orbital
Answer
(a) is a π bond with a node along the axis connecting the nuclei while (b) and (c) are σ bonds that overlap along the axis.
Thinking in terms of overlapping atomic orbitals is one way for us to explain how chemical bonds form in diatomic molecules.
However, to understand how molecules with more than two atoms form stable bonds, we require a more detailed model. As an
example, let us consider the water molecule, in which we have one oxygen atom bonding to two hydrogen atoms. Oxygen has the
electron configuration 1s22s22p4, with two unpaired electrons (one in each of the two 2p orbitals). Valence bond theory would
predict that the two O–H bonds form from the overlap of these two 2p orbitals with the 1s orbitals of the hydrogen atoms. If this
were the case, the bond angle would be 90°, as shown in Figure 5.3.5, because p orbitals are perpendicular to each other.
Experimental evidence shows that the bond angle is 104.5°, not 90°. The prediction of the valence bond theory model does not
match the real-world observations of a water molecule; a different model is needed.
Figure 5.3.5 : The hypothetical overlap of two of the 2p orbitals on an oxygen atom (red) with the 1s orbitals of two hydrogen
atoms (blue) would produce a bond angle of 90°. This is not consistent with experimental evidence.1
Quantum-mechanical calculations suggest why the observed bond angles in H2O differ from those predicted by the overlap of the
1s orbital of the hydrogen atoms with the 2p orbitals of the oxygen atom. The mathematical expression known as the wave
function, ψ, contains information about each orbital and the wavelike properties of electrons in an isolated atom. When atoms are
5.3.4 https://chem.libretexts.org/@go/page/126655
bound together in a molecule, the wave functions combine to produce new mathematical descriptions that have different shapes.
This process of combining the wave functions for atomic orbitals is called hybridization and is mathematically accomplished by the
linear combination of atomic orbitals, LCAO, (a technique that we will encounter again later). The new orbitals that result are
called hybrid orbitals. The valence orbitals in an isolated oxygen atom are a 2s orbital and three 2p orbitals. The valence orbitals in
an oxygen atom in a water molecule differ; they consist of four equivalent hybrid orbitals that point approximately toward the
corners of a tetrahedron (Figure 5.3.6). Consequently, the overlap of the O and H orbitals should result in a tetrahedral bond angle
(109.5°). The observed angle of 104.5° is experimental evidence for which quantum-mechanical calculations give a useful
explanation: Valence bond theory must include a hybridization component to give accurate predictions.
Figure 5.3.6 : (a) A water molecule has four regions of electron density, so VSEPR theory predicts a tetrahedral arrangement of
hybrid orbitals. (b) Two of the hybrid orbitals on oxygen contain lone pairs, and the other two overlap with the 1s orbitals of
hydrogen atoms to form the O–H bonds in H2O. This description is more consistent with the experimental structure.
The following ideas are important in understanding hybridization:
1. Hybrid orbitals do not exist in isolated atoms. They are formed only in covalently bonded atoms.
2. Hybrid orbitals have shapes and orientations that are very different from those of the atomic orbitals in isolated atoms.
3. A set of hybrid orbitals is generated by combining atomic orbitals. The number of hybrid orbitals in a set is equal to the number
of atomic orbitals that were combined to produce the set.
4. All orbitals in a set of hybrid orbitals are equivalent in shape and energy.
5. The type of hybrid orbitals formed in a bonded atom depends on its electron-pair geometry as predicted by the VSEPR theory.
6. Hybrid orbitals overlap to form σ bonds. Unhybridized orbitals overlap to form π bonds.
In the following sections, we shall discuss the common types of hybrid orbitals.
sp Hybridization
The beryllium atom in a gaseous BeCl2 molecule is an example of a central atom with no lone pairs of electrons in a linear
arrangement of three atoms. There are two regions of valence electron density in the BeCl2 molecule that correspond to the two
covalent Be–Cl bonds. To accommodate these two electron domains, two of the Be atom’s four valence orbitals will mix to yield
two hybrid orbitals. This hybridization process involves mixing of the valence s orbital with one of the valence p orbitals to yield
two equivalent sp hybrid orbitals that are oriented in a linear geometry (Figure 5.3.7). In this figure, the set of sp orbitals appears
similar in shape to the original p orbital, but there is an important difference. The number of atomic orbitals combined always
equals the number of hybrid orbitals formed. The p orbital is one orbital that can hold up to two electrons. The sp set is two
equivalent orbitals that point 180° from each other. The two electrons that were originally in the s orbital are now distributed to the
two sp orbitals, which are half filled. In gaseous BeCl2, these half-filled hybrid orbitals will overlap with orbitals from the chlorine
atoms to form two identical σ bonds.
5.3.5 https://chem.libretexts.org/@go/page/126655
Figure 5.3.7 : Hybridization of an s orbital (blue) and a p orbital (red) of the same atom produces two sp hybrid orbitals (yellow).
Each hybrid orbital is oriented primarily in just one direction. Note that each sp orbital contains one lobe that is significantly
larger than the other. The set of two sp orbitals are oriented at 180°, which is consistent with the geometry for two domains.
We illustrate the electronic differences in an isolated Be atom and in the bonded Be atom in the orbital energy-level diagram in
Figure 5.3.8. These diagrams represent each orbital by a horizontal line (indicating its energy) and each electron by an arrow.
Energy increases toward the top of the diagram. We use one upward arrow to indicate one electron in an orbital and two arrows (up
and down) to indicate two electrons of opposite spin.
Figure 5.3.8 : This orbital energy-level diagram shows the sp hybridized orbitals on Be in the linear BeCl2 molecule. Each of the
two sp hybrid orbitals holds one electron and is thus half filled and available for bonding via overlap with a Cl 3p orbital.
When atomic orbitals hybridize, the valence electrons occupy the newly created orbitals. The Be atom had two valence electrons,
so each of the sp orbitals gets one of these electrons. Each of these electrons pairs up with the unpaired electron on a chlorine atom
when a hybrid orbital and a chlorine orbital overlap during the formation of the Be–Cl bonds. Any central atom surrounded by just
two regions of valence electron density in a molecule will exhibit sp hybridization. Other examples include the mercury atom in the
linear HgCl2 molecule, the zinc atom in Zn(CH3)2, which contains a linear C–Zn–C arrangement, and the carbon atoms in HCCH
and CO2.
sp2 Hybridization
The valence orbitals of a central atom surrounded by three regions of electron density consist of a set of three sp2 hybrid orbitals
and one unhybridized p orbital. This arrangement results from sp2 hybridization, the mixing of one s orbital and two p orbitals to
produce three identical hybrid orbitals oriented in a trigonal planar geometry (Figure 5.3.9).
5.3.6 https://chem.libretexts.org/@go/page/126655
Figure 5.3.9 : The hybridization of an s orbital (blue) and two p orbitals (red) produces three equivalent sp2 hybridized orbitals
(yellow) oriented at 120° with respect to each other. The remaining unhybridized p orbital is not shown here, but is located along
the z axis.
Although quantum mechanics yields the “plump” orbital lobes as depicted in Figure 5.3.5, sometimes for clarity these orbitals are
drawn thinner and without the minor lobes, as in Figure 5.3.10, to avoid obscuring other features of a given illustration. We will
use these “thinner” representations whenever the true view is too crowded to easily visualize.
Figure 5.3.10 : This alternate way of drawing the trigonal planar sp2 hybrid orbitals is sometimes used in more crowded figures.
The observed structure of the borane molecule, BH3, suggests sp2 hybridization for boron in this compound. The molecule is
trigonal planar, and the boron atom is involved in three bonds to hydrogen atoms ( Figure 5.3.11).
5.3.7 https://chem.libretexts.org/@go/page/126655
three sp2 hybrid orbitals, and each boron electron pairs with a hydrogen electron when B–H bonds form.
Figure 5.3.12 : In an isolated B atom, there are one 2s and three 2p valence orbitals. When boron is in a molecule with three
regions of electron density, three of the orbitals hybridize and create a set of three sp2 orbitals and one unhybridized 2p orbital. The
three half-filled hybrid orbitals each overlap with an orbital from a hydrogen atom to form three σ bonds in BH3.
Any central atom surrounded by three regions of electron density will exhibit sp2 hybridization. This includes molecules with a
lone pair on the central atom, such as ClNO (Figure 5.3.13), or molecules with two single bonds and a double bond connected to
the central atom, as in formaldehyde, CH2O, and ethene, H2CCH2.
Figure 5.3.13 : The central atom(s) in each of the structures shown contain three regions of electron density and are sp2
hybridized. As we know from the discussion of VSEPR theory, a region of electron density contains all of the electrons that point in
one direction. A lone pair, an unpaired electron, a single bond, or a multiple bond would each count as one region of electron
density.
sp3 Hybridization
The valence orbitals of an atom surrounded by a tetrahedral arrangement of bonding pairs and lone pairs consist of a set of four sp3
hybrid orbitals. The hybrids result from the mixing of one s orbital and all three p orbitals that produces four identical sp3 hybrid
orbitals (Figure 5.3.14). Each of these hybrid orbitals points toward a different corner of a tetrahedron.
5.3.8 https://chem.libretexts.org/@go/page/126655
Figure 5.3.14 : The hybridization of an s orbital (blue) and three p orbitals (red) produces four equivalent sp3 hybridized orbitals
(yellow) oriented at 109.5° with respect to each other.
A molecule of methane, CH4, consists of a carbon atom surrounded by four hydrogen atoms at the corners of a tetrahedron. The
carbon atom in methane exhibits sp3 hybridization. We illustrate the orbitals and electron distribution in an isolated carbon atom
and in the bonded atom in CH4 in Figure 5.3.15. The four valence electrons of the carbon atom are distributed equally in the hybrid
orbitals, and each carbon electron pairs with a hydrogen electron when the C–H bonds form.
Figure 5.3.15 : The four valence atomic orbitals from an isolated carbon atom all hybridize when the carbon bonds in a molecule
like CH4 with four regions of electron density. This creates four equivalent sp3 hybridized orbitals. Overlap of each of the hybrid
orbitals with a hydrogen orbital creates a C–H σ bond.
In a methane molecule, the 1s orbital of each of the four hydrogen atoms overlaps with one of the four sp3 orbitals of the carbon
atom to form a sigma (σ) bond. This results in the formation of four strong, equivalent covalent bonds between the carbon atom and
each of the hydrogen atoms to produce the methane molecule, CH4.
The structure of ethane, C2H6, is similar to that of methane in that each carbon in ethane has four neighboring atoms arranged at the
corners of a tetrahedron—three hydrogen atoms and one carbon atom (Figure 5.3.10). However, in ethane an sp3 orbital of one
5.3.9 https://chem.libretexts.org/@go/page/126655
carbon atom overlaps end to end with an sp3 orbital of a second carbon atom to form a σ bond between the two carbon atoms. Each
of the remaining sp3 hybrid orbitals overlaps with an s orbital of a hydrogen atom to form carbon–hydrogen σ bonds. The structure
and overall outline of the bonding orbitals of ethane are shown in Figure 5.3.16. The orientation of the two CH3 groups is not fixed
relative to each other. Experimental evidence shows that rotation around σ bonds occurs easily.
Figure 5.3.16 : (a) In the ethane molecule, C2H6, each carbon has four sp3 orbitals. (b) These four orbitals overlap to form seven
σ bonds.
An sp3 hybrid orbital can also hold a lone pair of electrons. For example, the nitrogen atom in ammonia is surrounded by three
bonding pairs and a lone pair of electrons directed to the four corners of a tetrahedron. The nitrogen atom is sp3 hybridized with
one hybrid orbital occupied by the lone pair.
The molecular structure of water is consistent with a tetrahedral arrangement of two lone pairs and two bonding pairs of electrons.
Thus we say that the oxygen atom is sp3 hybridized, with two of the hybrid orbitals occupied by lone pairs and two by bonding
pairs. Since lone pairs occupy more space than bonding pairs, structures that contain lone pairs have bond angles slightly distorted
from the ideal. Perfect tetrahedra have angles of 109.5°, but the observed angles in ammonia (107.3°) and water (104.5°) are
slightly smaller. Other examples of sp3 hybridization include CCl4, PCl3, and NCl3.
5.3.10 https://chem.libretexts.org/@go/page/126655
Figure 5.3.17 : The shapes of hybridized orbital sets are consistent with the electron-pair geometries. For example, an atom
surrounded by three regions of electron density is sp2 hybridized, and the three sp2 orbitals are arranged in a trigonal planar fashion.
For example, we have discussed the H–O–H bond angle in H2O, 104.5°, which is more consistent with sp3 hybrid orbitals (109.5°)
on the central atom than with 2p orbitals (90°). Sulfur is in the same group as oxygen, and H2S has a similar Lewis structure.
However, it has a much smaller bond angle (92.1°), which indicates much less hybridization on sulfur than oxygen. Continuing
down the group, tellurium is even larger than sulfur, and for H2Te, the observed bond angle (90°) is consistent with overlap of the
5p orbitals, without invoking hybridization. We invoke hybridization where it is necessary to explain the observed structures.
5.3.11 https://chem.libretexts.org/@go/page/126655
The hybridization is sp3. since the carbon has a tetrahedral geometry (Figure 5.3.17).
The nitrogen atoms are surrounded by four regions of electron density, which arrange themselves in a tetrahedral electron-pair
geometry. The hybridization in a tetrahedral arrangement is sp3 (Figure 5.3.17). This is the hybridization of the nitrogen atoms
in urea.
The carbon atom is surrounded by three regions of electron density, positioned in a trigonal planar arrangement. The
hybridization in a trigonal planar electron pair geometry is sp2 (Figure 5.3.17), which is the hybridization of the carbon atom in
urea.
Exercise 5.3.2
Acetic acid, H3CC(O)OH, is the molecule that gives vinegar its odor and sour taste. What is the hybridization of the two carbon
atoms in acetic acid?
Answer
H3C, sp3; C(O)OH, sp2
Summary
5.3.12 https://chem.libretexts.org/@go/page/126655
The Molecular Shape of You (Ed Sheera…
Sheera…
Video 5.3.2 : An overview of electron involvment in bonding. Good luck getting this tune out of your head.
Valence bond theory describes bonding as a consequence of the overlap of two separate atomic orbitals on different atoms that
creates a region with one pair of electrons shared between the two atoms. When the orbitals overlap along an axis containing the
nuclei, they form a σ bond. When they overlap in a fashion that creates a node along this axis, they form a π bond.
We can use hybrid orbitals, which are mathematical combinations of some or all of the valence atomic orbitals, to describe the
electron density around covalently bonded atoms. These hybrid orbitals either form sigma (σ) bonds directed toward other atoms of
the molecule or contain lone pairs of electrons. We can determine the type of hybridization around a central atom from the
geometry of the regions of electron density about it. Two such regions imply sp hybridization; three, sp2 hybridization; four, sp3
hybridization. Pi (π) bonds are formed from unhybridized atomic orbitals (p or d orbitals).
Footnotes
1. Note that orbitals may sometimes be drawn in an elongated “balloon” shape rather than in a more realistic “plump” shape in
order to make the geometry easier to visualize.
Glossary
hybrid orbital
orbital created by combining atomic orbitals on a central atom
hybridization
model that describes the changes in the atomic orbitals of an atom when it forms a covalent compound
overlap
coexistence of orbitals from two different atoms sharing the same region of space, leading to the formation of a covalent bond
node
plane separating different lobes of orbitals, where the probability of finding an electron is zero
pi bond (π bond)
covalent bond formed by side-by-side overlap of atomic orbitals; the electron density is found on opposite sides of the
internuclear axis
sigma bond (σ bond)
covalent bond formed by overlap of atomic orbitals along the internuclear axis
valence bond theory
description of bonding that involves atomic orbitals overlapping to form σ or π bonds, within which pairs of electrons are
shared
sp hybrid orbital
one of a set of two orbitals with a linear arrangement that results from combining one s and one p orbital
5.3.13 https://chem.libretexts.org/@go/page/126655
one of a set of three orbitals with a trigonal planar arrangement that results from combining one s and two p orbitals
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
A Capella Science is a production of Tim Blaise.
Feedback
5.3: Valence Bond Theory and Hybrid Orbitals is shared under a CC BY license and was authored, remixed, and/or curated by LibreTexts.
5.3.14 https://chem.libretexts.org/@go/page/126655
5.3: Valence Bond Theory and Hybrid Orbitals (Problems)
PROBLEM 5.3.1
Explain how σ and π bonds are similar and how they are different.
Answer
Similarities: Both types of bonds result from overlap of atomic orbitals on adjacent atoms and contain a maximum of two
electrons. Differences: σ bonds are stronger and result from end-to-end overlap and all single bonds are σ bonds; π bonds
between the same two atoms are weaker because they result from side-by-side overlap, and multiple bonds contain one or
more π bonds (in addition to a σ bond).
PROBLEM 5.3.2
How many σ and π bonds are present in the molecule HCN?
Answer
H– C ≡ N has two σ (H–C and C–N) and two π (making the CN triple bond).
PROBLEM 5.3.3
A friend tells you N2 has three π bonds due to overlap of the three p-orbitals on each N atom. Do you agree?
Answer
No, two of the p orbitals (one on each N) will be oriented end-to-end and will form a σ bond.
PROBLEM 5.3.4
Draw the Lewis structures for CO2 and CO, and predict the number of σ and π bonds for each molecule.
a. CO2
b. CO
Answer a
2σ2π
Answer b
1σ2π
PROBLEM 5.3.5
5.3.1 https://chem.libretexts.org/@go/page/126656
Why is the concept of hybridization required in valence bond theory?
Answer
Hybridization is introduced to explain the geometry of bonding orbitals in valance bond theory.
PROBLEM 5.3.6
Give the shape that describes each hybrid orbital set:
(a) sp2
(b) sp
Answer a
Answer b
PROBLEM 5.3.7
What is the hybridization of the central atom in each of the following?
5.3.2 https://chem.libretexts.org/@go/page/126656
(a) BeH2
(b) PO 3−
Answer a
linear
Answer b
tetrahedral
PROBLEM 5.3.8
A molecule with the formula AB3 could have one of two different shapes. Give the shape and the hybridization of the central A
atom for each.
Answer
trigonal planar, sp2; trigonal pyramidal (one lone pair on A) sp3
PROBLEM 5.3.9
Methionine, CH3SCH2CH2CH(NH2)CO2H, is an amino acid found in proteins. Based on the Lewis structure of this compound,
what is the hybridization type of each carbon, oxygen, the nitrogen, and the sulfur?
A Lewis structure is shown in which a carbon atom is single bonded to three hydrogen atoms and single bonded to a sulfur atom with two lone pairs of electrons. The sulfur atom is attached to a chain of four
singly bonded carbon atoms, the first two of which are single bonded to two hydrogen atoms each, and the third of which is single bonded to a hydrogen atom and single bonded to a nitrogen atom which has one
lone electron pair. The nitrogen atom is also single bonded to two hydrogen atoms. The fourth andfinal carbon in the chain is double bonded to an oxygen with two lone pairs of electrons and single bonded to an
oxygen atom with two lone pairs of electrons. The second oxygen atom is single bonded to a hydrogen atom.
Answer
Problem 5.3.9
5.3.3 https://chem.libretexts.org/@go/page/126656
PROBLEM 5.3.10
Two important industrial chemicals, ethene, C2H4, and propene, C3H6, are produced by the steam (or thermal) cracking process:
Answer a
Answer b
C3H8: tetrahedral
C2H4: trigonal planar
C3H6: trigonal planar (1&2) and tetrahedral (3)
CH4: tetrahedral
Answer c
5.3.4 https://chem.libretexts.org/@go/page/126656
Problem 5.3.10
PROBLEM 5.3.11
Consider nitrous acid, HNO2 (HONO).
(a) Write a Lewis structure.
(b) What are the electron pair and molecular geometries of the internal oxygen and nitrogen atoms in the HNO2 molecule?
(c) What is the hybridization on the internal oxygen and nitrogen atoms in HNO2?
Answer a
Answer b
Electron pair: O: tetrahedral, N: trigonal planar
Molecular geometry: O: bent (109), N: trigonal planar
Answer c
O: sp3
N: sp2
Problem 5.3.11
PROBLEM 5.3.12
5.3.5 https://chem.libretexts.org/@go/page/126656
Identify the hybridization of each carbon atom in the following molecule. (The arrangement of atoms is given; you need to
determine how many bonds connect each pair of atoms.)
A Lewis structure is shown that is missing all of its bonds. Six carbon atoms form a chain. There are three hydrogen atoms located around the first carbon, two located around the second, one located near the
fifth, and two located around the sixth carbon.
Answer
Problem 5.3.12
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Feedback
Think one of the answers above is wrong? Let us know here.
5.3: Valence Bond Theory and Hybrid Orbitals (Problems) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by LibreTexts.
5.3.6 https://chem.libretexts.org/@go/page/126656
CHAPTER OVERVIEW
Unit 6: Molecular Polarity
Unit Objectives
Define electronegativity and assess the polarity of covalent bonds
Explain the concepts of polar covalent bonds and molecular polarity
Assess the polarity of a molecule based on its bonding and structure
Describe the types of intermolecular forces possible between atoms or molecules in condensed phases (dispersion forces,
dipole-dipole attractions, and hydrogen bonding)
Identify the types of intermolecular forces experienced by specific molecules based on their structures
Explain the relation between the intermolecular forces present within a substance and the temperatures associated with
changes in its physical state
Unit 6: Molecular Polarity is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
1
6.1: Electronegativity and Polarity
Skills to Develop
Define electronegativity and assess the polarity of covalent bonds
Electronegativity Review
Whether a bond is nonpolar or polar covalent is determined by a property of the bonding atoms called electronegativity.
Electronegativity is a measure of the tendency of an atom to attract electrons (or electron density) towards itself. It determines how
the shared electrons are distributed between the two atoms in a bond. The more strongly an atom attracts the electrons in its bonds,
the larger its electronegativity. Electrons in a polar covalent bond are shifted toward the more electronegative atom; thus, the more
electronegative atom is the one with the partial negative charge. The greater the difference in electronegativity, the more polarized
the electron distribution and the larger the partial charges of the atoms.
Figure 6.1.1 shows the electronegativity values of the elements as proposed by one of the most famous chemists of the twentieth
century: Linus Pauling. In general, electronegativity increases from left to right across a period in the periodic table and decreases
down a group. Thus, the nonmetals, which lie in the upper right, tend to have the highest electronegativities, with fluorine the most
electronegative element of all (EN = 4.0). Metals tend to be less electronegative elements, and the group 1 metals have the lowest
electronegativities. Note that noble gases are excluded from this figure because these atoms usually do not share electrons with
others atoms since they have a full valence shell. (While noble gas compounds such as XeO2 do exist, they can only be formed
under extreme conditions, and thus they do not fit neatly into the general model of electronegativity.)
Figure 6.1.1 : The electronegativity values derived by Pauling follow predictable periodic trends with the higher
electronegativities toward the upper right of the periodic table.
Linus Pauling
6.1.1 https://chem.libretexts.org/@go/page/112692
Linus Pauling is the only person to have received two unshared (individual) Nobel Prizes: one for chemistry in 1954 for his
work on the nature of chemical bonds and one for peace in 1962 for his opposition to weapons of mass destruction. He
developed many of the theories and concepts that are foundational to our current understanding of chemistry, including
electronegativity and resonance structures.
Linus Pauling (1901–1994) made many important contributions to the field of chemistry. He was also a prominent activist,
publicizing issues related to health and nuclear weapons.
Pauling also contributed to many other fields besides chemistry. His research on sickle cell anemia revealed the cause of the
disease—the presence of a genetically inherited abnormal protein in the blood—and paved the way for the field of molecular
genetics. His work was also pivotal in curbing the testing of nuclear weapons; he proved that radioactive fallout from nuclear
testing posed a public health risk.
Figure 6.1.2 : As the electronegativity difference increases between two atoms, the bond becomes more ionic.
A rough approximation of the electronegativity differences associated with covalent, polar covalent, and ionic bonds is shown in
Figure 6.1.4. This table is just a general guide, however, with many exceptions. For example, the H and F atoms in HF have an
electronegativity difference of 1.9, and the N and H atoms in NH3 a difference of 0.9, yet both of these compounds form bonds that
are considered polar covalent. Likewise, the Na and Cl atoms in NaCl have an electronegativity difference of 2.1, and the Mn and I
atoms in MnI2 have a difference of 1.0, yet both of these substances form ionic compounds.
6.1.2 https://chem.libretexts.org/@go/page/112692
The best guide to the covalent or ionic character of a bond is to consider the types of atoms involved and their relative positions in
the periodic table. Bonds between two nonmetals are generally covalent; bonding between a metal and a nonmetal is often ionic.
Some compounds contain both covalent and ionic bonds. The atoms in polyatomic ions, such as OH–, NO , and NH , are held
−
3
+
4
together by polar covalent bonds. However, these polyatomic ions form ionic compounds by combining with ions of opposite
charge. For example, potassium nitrate, KNO3, contains the K+ cation and the polyatomic NO anion. Thus, bonding in potassium
−
3
nitrate is ionic, resulting from the electrostatic attraction between the ions K+ and NO , as well as covalent between the nitrogen
−
3
designate the positive and negative atoms using the symbols δ+ and δ–:
C–H, C–N, C–O, N–H, O–H, S–H
Solution
The polarity of these bonds increases as the absolute value of the electronegativity difference increases. The atom with the δ–
designation is the more electronegative of the two. Table 6.1.1 shows these bonds in order of increasing polarity.
Table 6.1.1: Bond Polarity and Electronegativity Difference
Bond ΔEN Polarity
δ− δ+
C–H 0.4 C − H
δ− δ+
S–H 0.4 S − H
δ+ δ−
C–N 0.5 C − N
δ− δ+
N–H 0.9 N − H
δ+ δ−
C–O 1.0 C − O
δ− δ+
O–H 1.4 O − H
Exercise 6.1.1
Silicones are polymeric compounds containing, among others, the following types of covalent bonds: Si–O, Si–C, C–H, and C–
C. Using the electronegativity values in Figure 6.1.3, arrange the bonds in order of increasing polarity and designate the positive
and negative atoms using the symbols δ+ and δ–.
Answer
δ+ δ−
Si–C 0.7 Si − C
δ+ δ−
Si–O 1.7 Si − O
Learn More
6.1.3 https://chem.libretexts.org/@go/page/112692
How polarity makes water behave stra…
stra…
Summary
Electronegativity
Glossary
bond length
distance between the nuclei of two bonded atoms at which the lowest potential energy is achieved
covalent bond
bond formed when electrons are shared between atoms
electronegativity
tendency of an atom to attract electrons in a bond to itself
6.1.4 https://chem.libretexts.org/@go/page/112692
(also, nonpolar covalent bond) covalent bond between atoms of identical electronegativities
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
TED-Ed’s commitment to creating lessons worth sharing is an extension of TED’s mission of spreading great ideas. Within
TED-Ed’s growing library of TED-Ed animations, you will find carefully curated educational videos, many of which represent
collaborations between talented educators and animators nominated through the TED-Ed website.
Teacher's Pet
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
6.1: Electronegativity and Polarity is shared under a CC BY license and was authored, remixed, and/or curated by LibreTexts.
6.1.5 https://chem.libretexts.org/@go/page/112692
6.1: Electronegativity and Polarity (Problems)
PROBLEM 6.1.1
Predict which of the following compounds are ionic and which are covalent, based on the location of their constituent atoms in
the periodic table:
a. Cl2CO
b. MnO
c. NCl3
d. CoBr2
e. K2S
f. CO
g. CaF2
h. HI
i. CaO
j. IBr
k. CO2
Answer a
covalent
Answer b
ionic
Answer c
covalent
Answer d
ionic
Answer e
ionic
Answer f
covalent
Answer g
ionic
Answer h
covalent
Answer i
ionic
Answer j
covalent
Answer k
covalent
PROBLEM 6.1.2
6.1.1 https://chem.libretexts.org/@go/page/112694
Explain the difference between a nonpolar covalent bond, a polar covalent bond, and an ionic bond.
Answer
nonpolar covalent: electronegativity difference is less than 0.4 (nonmetal+nonmetal close together on the periodic table)
polar covalent: electronegativity difference in between 0.4 and 2.0 (nonmetal + nonmental further apart on the periodic table)
ionic: electronegativity difference is above 2.0 (metal + nonmetal)
PROBLEM 6.1.3
From its position in the periodic table, determine which atom in each pair is more electronegative:
a. Br or Cl
b. N or O
c. S or O
d. P or S
e. Si or N
f. Ba or P
g. N or K
Answer a
Cl
Answer b
O
Answer c
O
Answer d
S
Answer e
N
Answer f
P
Answer g
N
PROBLEM 6.1.4
From its position in the periodic table, determine which atom in each pair is more electronegative:
a. N or P
b. N or Ge
c. S or F
d. Cl or S
e. H or C
f. Se or P
g. C or Si
Answer a
6.1.2 https://chem.libretexts.org/@go/page/112694
N
Answer b
N
Answer c
F
Answer d
Cl
Answer e
C
Answer f
P
Answer g
C
PROBLEM 6.1.5
From their positions in the periodic table, arrange the atoms in each of the following series in order of increasing
electronegativity:
a. C, F, H, N, O
b. Br, Cl, F, H, I
c. F, H, O, P, S
d. Al, H, Na, O, P
e. Ba, H, N, O, As
Answer a
H, C, N, O, F
Answer b
H, I, Br, Cl, F
Answer c
H, P, S, O, F
Answer d
Na, Al, H, P, O
Answer e
Ba, H, As, N, O
PROBLEM 6.1.6
Which is the most polar bond?
C–C; C–H; N–H; O–H; Se–H
Answer
O–H
6.1.3 https://chem.libretexts.org/@go/page/112694
PROBLEM 6.1.7
Identify the more polar bond in each of the following pairs of bonds:
a. HF or HCl
b. NO or CO
c. SH or OH
d. PCl or SCl
e. CH or NH
f. SO or PO
g. CN or NN
Answer a
HF
Answer b
CO
Answer c
OH
Answer d
PCl
Answer e
NH
Answer f
PO
Answer g
CN
PROBLEM 6.1.8
Which of the following molecules or ions contain polar bonds?
a. O3
b. O2−
2
c. NO −
3
d. CO2
e. H2S
f. BH −
g. S8
Answer
c, d, e, f
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
6.1.4 https://chem.libretexts.org/@go/page/112694
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Feedback
Think one of the answers above is wrong? Let us know here.
6.1: Electronegativity and Polarity (Problems) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.
6.1.5 https://chem.libretexts.org/@go/page/112694
6.2: Molecular Shape and Polarity
Skills to Develop
Explain the concepts of polar covalent bonds and molecular polarity
Assess the polarity of a molecule based on its bonding and structure
μ = Qr (6.2.1)
where
Q is the magnitude of the partial charges (determined by the electronegativity difference) and
r is the distance between the charges:
This bond moment can be represented as a vector, a quantity having both direction and magnitude (Figure 6.2.1). Dipole vectors are shown as arrows
pointing along the bond from the less electronegative atom toward the more electronegative atom. A small plus sign is drawn on the less electronegative
end to indicate the partially positive end of the bond. The length of the arrow is proportional to the magnitude of the electronegativity difference between
the two atoms.
Figure 6.2.1 : (a) There is a small difference in electronegativity between C and H, represented as a short vector. (b) The electronegativity difference
between B and F is much larger, so the vector representing the bond moment is much longer.
A whole molecule may also have a separation of charge, depending on its molecular structure and the polarity of each of its bonds. If such a charge
separation exists, the molecule is said to be a polar molecule (or dipole); otherwise the molecule is said to be nonpolar. The dipole moment measures the
extent of net charge separation in the molecule as a whole. We determine the dipole moment by adding the bond moments in three-dimensional space,
taking into account the molecular structure.
For diatomic molecules, there is only one bond, so its bond dipole moment determines the molecular polarity. Homonuclear diatomic molecules such as Br2
and N2 have no difference in electronegativity, so their dipole moment is zero. For heteronuclear molecules such as CO, there is a small dipole moment. For
HF, there is a larger dipole moment because there is a larger difference in electronegativity.
When a molecule contains more than one bond, the geometry must be taken into account. If the bonds in a molecule are arranged such that their bond
moments cancel (vector sum equals zero), then the molecule is nonpolar. This is the situation in CO2 (Figure 6.2.2A). Each of the bonds is polar, but the
molecule as a whole is nonpolar. From the Lewis structure, and using VSEPR theory, we determine that the CO2 molecule is linear with polar C=O bonds
on opposite sides of the carbon atom. The bond moments cancel because they are pointed in opposite directions. In the case of the water molecule (Figure
6.2.2B), the Lewis structure again shows that there are two bonds to a central atom, and the electronegativity difference again shows that each of these
bonds has a nonzero bond moment. In this case, however, the molecular structure is bent because of the lone pairs on O, and the two bond moments do not
cancel. Therefore, water does have a net dipole moment and is a polar molecule (dipole).
Figure 6.2.2 : The overall dipole moment of a molecule depends on the individual bond dipole moments and how they are arranged. (a) Each CO bond has
a bond dipole moment, but they point in opposite directions so that the net CO2 molecule is nonpolar. (b) In contrast, water is polar because the OH bond
moments do not cancel out.
The OCS molecule has a structure similar to CO2, but a sulfur atom has replaced one of the oxygen atoms. To determine if this molecule is polar, we draw
the molecular structure. VSEPR theory predicts a linear molecule:
6.2.1 https://chem.libretexts.org/@go/page/112695
The C–O bond is considerably polar. Although C and S have very similar electronegativity values, S is slightly more electronegative than C, and so the C-S
bond is just slightly polar. Because oxygen is more electronegative than sulfur, the oxygen end of the molecule is the negative end.
Chloromethane, CH3Cl, is another example of a polar molecule. Although the polar C–Cl and C–H bonds are arranged in a tetrahedral geometry, the C–Cl
bonds have a larger bond moment than the C–H bond, and the bond moments do not completely cancel each other. All of the dipoles have a upward
component in the orientation shown, since carbon is more electronegative than hydrogen and less electronegative than chlorine:
When we examine the highly symmetrical molecules BF3 (trigonal planar), CH4 (tetrahedral), PF5 (trigonal bipyramidal), and SF6 (octahedral), in which all
the polar bonds are identical, the molecules are nonpolar. The bonds in these molecules are arranged such that their dipoles cancel. However, just because a
molecule contains identical bonds does not mean that the dipoles will always cancel. Many molecules that have identical bonds and lone pairs on the central
atoms have bond dipoles that do not cancel. Examples include H2S and NH3. A hydrogen atom is at the positive end and a nitrogen or sulfur atom is at the
negative end of the polar bonds in these molecules:
Figure 6.2.3 : (a) Molecules are always randomly distributed in the liquid state in the absence of an electric field. (b) When an electric field is applied,
polar molecules like HF will align to the dipoles with the field direction.
6.2.2 https://chem.libretexts.org/@go/page/112695
Molecule Polarity
B
A C
A B
Three Atoms Real Molecules
Two Atoms
Using the above simulator, select the “Three Atoms” tab at the top. This should display a molecule ABC with three electronegativity adjustors. You can
display or hide the bond moments, molecular dipoles, and partial charges at the right. Turning on the Electric Field will show whether the molecule
moves when exposed to a field, similar to Figure 6.2.14.
Use the electronegativity controls to determine how the molecular dipole will look for the starting bent molecule if:
a. A and C are very electronegative and B is in the middle of the range.
b. A is very electronegative, and B and C are not.
Solution
a. Molecular dipole moment points immediately between A and C.
b. Molecular dipole moment points along the A–B bond, toward A.
Exercise 6.2.1
Determine the partial charges that will give the largest possible bond dipoles.
Answer
The largest bond moments will occur with the largest partial charges. The two solutions above represent how unevenly the electrons are shared in the
bond. The bond moments will be maximized when the electronegativity difference is greatest. The controls for A and C should be set to one extreme,
and B should be set to the opposite extreme. Although the magnitude of the bond moment will not change based on whether B is the most
electronegative or the least, the direction of the bond moment will.
Summary
6.2.3 https://chem.libretexts.org/@go/page/112695
Polar & Non-Polar Molecules: Crash Course Chem…
Chem…
Glossary
axial position
location in a trigonal bipyramidal geometry in which there is another atom at a 180° angle and the equatorial positions are at a 90° angle
bond angle
angle between any two covalent bonds that share a common atom
bond distance
(also, bond length) distance between the nuclei of two bonded atoms
dipole moment
property of a molecule that describes the separation of charge determined by the sum of the individual bond moments based on the molecular structure
electron-pair geometry
arrangement around a central atom of all regions of electron density (bonds, lone pairs, or unpaired electrons)
equatorial position
one of the three positions in a trigonal bipyramidal geometry with 120° angles between them; the axial positions are located at a 90° angle
linear
shape in which two outside groups are placed on opposite sides of a central atom
molecular structure
structure that includes only the placement of the atoms in the molecule
octahedral
shape in which six outside groups are placed around a central atom such that a three-dimensional shape is generated with four groups forming a square
and the other two forming the apex of two pyramids, one above and one below the square plane
polar molecule
(also, dipole) molecule with an overall dipole moment
tetrahedral
shape in which four outside groups are placed around a central atom such that a three-dimensional shape is generated with four corners and 109.5°
angles between each pair and the central atom
trigonal bipyramidal
shape in which five outside groups are placed around a central atom such that three form a flat triangle with 120° angles between each pair and the
central atom, and the other two form the apex of two pyramids, one above and one below the triangular plane
6.2.4 https://chem.libretexts.org/@go/page/112695
trigonal planar
shape in which three outside groups are placed in a flat triangle around a central atom with 120° angles between each pair and the central atom
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley (Stephen F. Austin State
University) with contributing authors. Textbook content produced by OpenStax College is licensed under a Creative Commons Attribution License 4.0
license. Download for free at http://cnx.org/contents/[email protected]).
Adelaide Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
6.2: Molecular Shape and Polarity is shared under a CC BY license and was authored, remixed, and/or curated by LibreTexts.
6.2.5 https://chem.libretexts.org/@go/page/112695
6.2: Molecular Shape and Polarity (Problems)
PROBLEM 6.2.1
Explain how a molecule that contains polar bonds can be nonpolar.
Answer
As long as the polar bonds are compensated (for example. two identical atoms are found directly across the central atom
from one another), the molecule can be nonpolar.
PROBLEM 6.2.2
Which of the following molecules and ions contain polar bonds? Which of these molecules and ions have dipole moments?
a. ClF5
b. ClO −
2
c. TeCl 2−
4
d. PCl3
e. SeF4
f. PH −
2
g. XeF2
Answer
All of these molecules and ions contain polar bonds. Only ClF5, ClO , PCl3, SeF4, and PH have dipole moments.
−
2
−
PROBLEM 6.2.3
Which of the following molecules have dipole moments?
a. CS2
b. SeS2
c. CCl2F2
d. PCl3 (P is the central atom)
e. ClNO (N is the central atom)
Answer
SeS2, CCl2F2, PCl3, and ClNO all have dipole moments.
PROBLEM 6.2.4
The molecule XF3 has a dipole moment. Is X boron or phosphorus?
Answer
P
PROBLEM 6.2.5
The molecule XCl2 has a dipole moment. Is X beryllium or sulfur?
Answer
S
PROBLEM 6.2.6
Is the Cl2BBCl2 molecule polar or nonpolar?
6.2.1 https://chem.libretexts.org/@go/page/112708
Answer
nonpolar
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Feedback
Think one of the answers above is wrong? Let us know here.
6.2: Molecular Shape and Polarity (Problems) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.
6.2.2 https://chem.libretexts.org/@go/page/112708
6.3: Intermolecular Forces
Skills to Develop
Describe the types of intermolecular forces possible between atoms or molecules in condensed phases (dispersion forces,
dipole-dipole attractions, and hydrogen bonding)
Identify the types of intermolecular forces experienced by specific molecules based on their structures
Explain the relation between the intermolecular forces present within a substance and the temperatures associated with
changes in its physical state
The great distances between atoms and molecules in a gaseous phase, and the corresponding absence of any significant interactions
between them, allows for simple descriptions of many physical properties that are the same for all gases, regardless of their
chemical identities. As described last term when we explored gases, this situation changes at high pressures and low temperatures
—conditions that permit the atoms and molecules to interact to a much greater extent.
Figure 6.3.1 : Solid carbon dioxide (“dry ice”, left) sublimes vigorously when placed in a liquid (right), cooling the liquid and
generating a fog of condensed water vapor above the cylinder. (credit: modification of work by Paul Flowers)
In the liquid and solid states, these interactions are of considerable strength and play an important role in determining a number of
physical properties that do depend on the chemical identity of the substance. In this chapter, the nature of these interactions and
their effects on various physical properties of liquid and solid phases will be examined.
As was the case for gaseous substances, the kinetic molecular theory may be used to explain the behavior of solids and liquids. In
the following description, the term particle will be used to refer to an atom, molecule, or ion. Note that we will use the popular
phrase “intermolecular attraction” to refer to attractive forces between the particles of a substance, regardless of whether these
particles are molecules, atoms, or ions.
Consider these two aspects of the molecular-level environments in solid, liquid, and gaseous matter:
Particles in a solid are tightly packed together and often arranged in a regular pattern; in a liquid, they are close together with no
regular arrangement; in a gas, they are far apart with no regular arrangement.
Particles in a solid vibrate about fixed positions and do not generally move in relation to one another; in a liquid, they move
past each other but remain in essentially constant contact; in a gas, they move independently of one another except when they
collide.
The differences in the properties of a solid, liquid, or gas reflect the strengths of the attractive forces between the atoms, molecules,
or ions that make up each phase. The phase in which a substance exists depends on the relative extents of its intermolecular forces
(IMFs) and the kinetic energies (KE) of its molecules. IMFs are the various forces of attraction that may exist between the atoms
and molecules of a substance due to electrostatic phenomena, as will be detailed in this module. These forces serve to hold particles
close together, whereas the particles’ KE provides the energy required to overcome the attractive forces and thus increase the
distance between particles. Figure 6.3.2 illustrates how changes in physical state may be induced by changing the temperature,
hence, the average KE, of a given substance.
6.3.1 https://chem.libretexts.org/@go/page/112709
Figure 6.3.2 : Transitions between solid, liquid, and gaseous states of a substance occur when conditions of temperature or
pressure favor the associated changes in intermolecular forces. (Note: The space between particles in the gas phase is much
greater than shown.)
As an example of the processes depicted in this figure, consider a sample of water. When gaseous water is cooled sufficiently, the
attractions between H2O molecules will be capable of holding them together when they come into contact with each other; the gas
condenses, forming liquid H2O. For example, liquid water forms on the outside of a cold glass as the water vapor in the air is
cooled by the cold glass, as seen in Figure 6.3.2.
Figure 6.3.3 : Condensation forms when water vapor in the air is cooled enough to form liquid water, such as (a) on the outside of
a cold beverage glass or (b) in the form of fog. (credit a: modification of work by Jenny Downing; credit b: modification of work by
Cory Zanker)
We can also liquefy many gases by compressing them, if the temperature is not too high. The increased pressure brings the
molecules of a gas closer together, such that the attractions between the molecules become strong relative to their KE.
Consequently, they form liquids. Butane, C4H10, is the fuel used in disposable lighters and is a gas at standard temperature and
pressure. Inside the lighter’s fuel compartment, the butane is compressed to a pressure that results in its condensation to the liquid
state, as shown in Figure 6.3.4.
6.3.2 https://chem.libretexts.org/@go/page/112709
Figure 6.3.4 : Gaseous butane is compressed within the storage compartment of a disposable lighter, resulting in its condensation
to the liquid state. (credit: modification of work by “Sam-Cat”/Flickr)
Finally, if the temperature of a liquid becomes sufficiently low, or the pressure on the liquid becomes sufficiently high, the
molecules of the liquid no longer have enough KE to overcome the IMF between them, and a solid forms. A more thorough
discussion of these and other changes of state, or phase transitions, is provided in a later module of this chapter.
Figure 6.3.5 : Intramolecular forces keep a molecule intact. Intermolecular forces hold multiple molecules together and determine
many of a substance’s properties.
All of the attractive forces between neutral atoms and molecules are known as van der Waals forces, although they are usually
referred to more informally as intermolecular attraction. We will consider the various types of IMFs in the next three sections of
this module.
Dispersion Forces
One of the three van der Waals forces is present in all condensed phases, regardless of the nature of the atoms or molecules
composing the substance. This attractive force is called the London dispersion force in honor of German-born American physicist
Fritz London who, in 1928, first explained it. This force is often referred to as simply the dispersion force. Because the electrons of
an atom or molecule are in constant motion (or, alternatively, the electron’s location is subject to quantum-mechanical variability),
at any moment in time, an atom or molecule can develop a temporary, instantaneous dipole if its electrons are distributed
asymmetrically. The presence of this dipole can, in turn, distort the electrons of a neighboring atom or molecule, producing an
induced dipole. These two rapidly fluctuating, temporary dipoles thus result in a relatively weak electrostatic attraction between the
species—a so-called dispersion force like that illustrated in Figure 6.3.6.
Figure 6.3.6 : Dispersion forces result from the formation of temporary dipoles, as illustrated here for two nonpolar diatomic
molecules.
6.3.3 https://chem.libretexts.org/@go/page/112709
Dispersion forces that develop between atoms in different molecules can attract the two molecules to each other. The forces are
relatively weak, however, and become significant only when the molecules are very close. Larger and heavier atoms and molecules
exhibit stronger dispersion forces than do smaller and lighter atoms and molecules. F2 and Cl2 are gases at room temperature
(reflecting weaker attractive forces); Br2 is a liquid, and I2 is a solid (reflecting stronger attractive forces). Trends in observed
melting and boiling points for the halogens clearly demonstrate this effect, as seen in Table 6.3.1.
Table 6.3.1: Melting and Boiling Points of the Halogens
Halogen Molar Mass Atomic Radius Melting Point Boiling Point
fluorine, F2 38 g/mol 72 pm 53 K 85 K
The increase in melting and boiling points with increasing atomic/molecular size may be rationalized by considering how the
strength of dispersion forces is affected by the electronic structure of the atoms or molecules in the substance. In a larger atom, the
valence electrons are, on average, farther from the nuclei than in a smaller atom. Thus, they are less tightly held and can more
easily form the temporary dipoles that produce the attraction. The measure of how easy or difficult it is for another electrostatic
charge (for example, a nearby ion or polar molecule) to distort a molecule’s charge distribution (its electron cloud) is known as
polarizability. A molecule that has a charge cloud that is easily distorted is said to be very polarizable and will have large
dispersion forces; one with a charge cloud that is difficult to distort is not very polarizable and will have small dispersion forces.
Exercise 6.3.1
6.3.4 https://chem.libretexts.org/@go/page/112709
Order the following hydrocarbons from lowest to highest boiling point: C2H6, C3H8, and C4H10.
Answer
All of these compounds are nonpolar and only have London dispersion forces: the larger the molecule, the larger the
dispersion forces and the higher the boiling point. The ordering from lowest to highest boiling point is therefore
C2H6 < C3H8 < C4H10.
The shapes of molecules also affect the magnitudes of the dispersion forces between them. For example, boiling points for the
isomers n-pentane, isopentane, and neopentane (shown in Figure 6.3.7) are 36 °C, 27 °C, and 9.5 °C, respectively. Even though
these compounds are composed of molecules with the same chemical formula, C5H12, the difference in boiling points suggests that
dispersion forces in the liquid phase are different, being greatest for n-pentane and least for neopentane. The elongated shape of n-
pentane provides a greater surface area available for contact between molecules, resulting in correspondingly stronger dispersion
forces. The more compact shape of isopentane offers a smaller surface area available for intermolecular contact and, therefore,
weaker dispersion forces. Neopentane molecules are the most compact of the three, offering the least available surface area for
intermolecular contact and, hence, the weakest dispersion forces. This behavior is analogous to the connections that may be formed
between strips of VELCRO brand fasteners: the greater the area of the strip’s contact, the stronger the connection.
Figure 6.3.7 : The strength of the dispersion forces increases with the contact area between molecules, as demonstrated by the
boiling points of these pentane isomers.
Applications: Geckos and Intermolecular Forces
Geckos have an amazing ability to adhere to most surfaces. They can quickly run up smooth walls and across ceilings that have
no toe-holds, and they do this without having suction cups or a sticky substance on their toes. And while a gecko can lift its feet
easily as it walks along a surface, if you attempt to pick it up, it sticks to the surface. How are geckos (as well as spiders and
some other insects) able to do this? Although this phenomenon has been investigated for hundreds of years, scientists only
recently uncovered the details of the process that allows geckos’ feet to behave this way.
6.3.5 https://chem.libretexts.org/@go/page/112709
Figure 6.3.7 : Geckos’ toes contain large numbers of tiny hairs (setae), which branch into many triangular tips (spatulae).
Geckos adhere to surfaces because of van der Waals attractions between the surface and a gecko’s millions of spatulae. By
changing how the spatulae contact the surface, geckos can turn their stickiness “on” and “off.” (credit photo: modification of
work by “JC*+A!”/Flickr)
Geckos’ toes are covered with hundreds of thousands of tiny hairs known as setae, with each seta, in turn, branching into
hundreds of tiny, flat, triangular tips called spatulae. The huge numbers of spatulae on its setae provide a gecko, shown in
Figure 6.3.7, with a large total surface area for sticking to a surface. In 2000, Kellar Autumn, who leads a multi-institutional
gecko research team, found that geckos adhered equally well to both polar silicon dioxide and nonpolar gallium arsenide. This
proved that geckos stick to surfaces because of dispersion forces—weak intermolecular attractions arising from temporary,
synchronized charge distributions between adjacent molecules. Although dispersion forces are very weak, the total attraction
over millions of spatulae is large enough to support many times the gecko’s weight.
In 2014, two scientists developed a model to explain how geckos can rapidly transition from “sticky” to “non-sticky.” Alex
Greaney and Congcong Hu at Oregon State University described how geckos can achieve this by changing the angle between
their spatulae and the surface. Geckos’ feet, which are normally nonsticky, become sticky when a small shear force is applied.
By curling and uncurling their toes, geckos can alternate between sticking and unsticking from a surface, and thus easily move
across it. Further investigations may eventually lead to the development of better adhesives and other applications.
Video 6.3.1 : Watch this video to learn more about Kellar Autumn’s research that determined that van der Waals forces are
responsible for a gecko’s ability to cling and climb.
Dipole-Dipole Attractions
Recall from the chapter on chemical bonding and molecular geometry that polar molecules have a partial positive charge on one
side and a partial negative charge on the other side of the molecule—a separation of charge called a dipole. Consider a polar
molecule such as hydrogen chloride, HCl. In the HCl molecule, the more electronegative Cl atom bears the partial negative charge,
whereas the less electronegative H atom bears the partial positive charge. An attractive force between HCl molecules results from
the attraction between the positive end of one HCl molecule and the negative end of another. This attractive force is called a dipole-
dipole attraction—the electrostatic force between the partially positive end of one polar molecule and the partially negative end of
another, as illustrated in Figure 6.3.8.
6.3.6 https://chem.libretexts.org/@go/page/112709
Figure 6.3.8 : This image shows two arrangements of polar molecules, such as HCl, that allow an attraction between the partial
negative end of one molecule and the partial positive end of another.
The effect of a dipole-dipole attraction is apparent when we compare the properties of HCl molecules to nonpolar F2 molecules.
Both HCl and F2 consist of the same number of atoms and have approximately the same molecular mass. At a temperature of 150
K, molecules of both substances would have the same average KE. However, the dipole-dipole attractions between HCl molecules
are sufficient to cause them to “stick together” to form a liquid, whereas the relatively weaker dispersion forces between nonpolar
F2 molecules are not, and so this substance is gaseous at this temperature. The higher normal boiling point of HCl (188 K)
compared to F2 (85 K) is a reflection of the greater strength of dipole-dipole attractions between HCl molecules, compared to the
attractions between nonpolar F2 molecules. We will often use values such as boiling or freezing points, or enthalpies of
vaporization or fusion, as indicators of the relative strengths of IMFs of attraction present within different substances.
Exercise 6.3.2
Predict which will have the higher boiling point: ICl or Br . Explain your reasoning.
2
Answer
ICl. ICl and Br2 have similar masses (~160 amu) and therefore experience similar London dispersion forces. ICl is polar and
thus also exhibits dipole-dipole attractions; Br2 is nonpolar and does not. The relatively stronger dipole-dipole attractions
require more energy to overcome, so ICl will have the higher boiling point.
Hydrogen Bonding
Nitrosyl fluoride (ONF, molecular mass 49 amu) is a gas at room temperature. Water (H2O, molecular mass 18 amu) is a liquid,
even though it has a lower molecular mass. We clearly cannot attribute this difference between the two compounds to dispersion
forces. Both molecules have about the same shape and ONF is the heavier and larger molecule. It is, therefore, expected to
experience more significant dispersion forces. Additionally, we cannot attribute this difference in boiling points to differences in
the dipole moments of the molecules. Both molecules are polar and exhibit comparable dipole moments. The large difference
between the boiling points is due to a particularly strong dipole-dipole attraction that may occur when a molecule contains a
hydrogen atom bonded to a fluorine, oxygen, or nitrogen atom (the three most electronegative elements). The very large difference
in electronegativity between the H atom (2.1) and the atom to which it is bonded (4.0 for an F atom, 3.5 for an O atom, or 3.0 for a
N atom), combined with the very small size of a H atom and the relatively small sizes of F, O, or N atoms, leads to highly
concentrated partial charges with these atoms. Molecules with F-H, O-H, or N-H moieties are very strongly attracted to similar
moieties in nearby molecules, a particularly strong type of dipole-dipole attraction called hydrogen bonding. Examples of hydrogen
bonds include HF⋯HF, H2O⋯HOH, and H3N⋯HNH2, in which the hydrogen bonds are denoted by dots. Figure 6.3.9 illustrates
hydrogen bonding between water molecules.
6.3.7 https://chem.libretexts.org/@go/page/112709
Figure 6.3.9 : Water molecules participate in multiple hydrogen-bonding interactions with nearby water molecules.
Despite use of the word “bond,” keep in mind that hydrogen bonds are intermolecular attractive forces, not intramolecular
attractive forces (covalent bonds). Hydrogen bonds are much weaker than covalent bonds, only about 5 to 10% as strong, but are
generally much stronger than other dipole-dipole attractions and dispersion forces.
Hydrogen bonds have a pronounced effect on the properties of condensed phases (liquids and solids). For example, consider the
trends in boiling points for the binary hydrides of group 15 (NH3, PH3, AsH3, and SbH3), group 16 hydrides (H2O, H2S, H2Se, and
H2Te), and group 17 hydrides (HF, HCl, HBr, and HI). The boiling points of the heaviest three hydrides for each group are plotted
in Figure 6.3.10. As we progress down any of these groups, the polarities of the molecules decrease slightly, whereas the sizes of
the molecules increase substantially. The effect of increasingly stronger dispersion forces dominates that of increasingly weaker
dipole-dipole attractions, and the boiling points are observed to increase steadily.
Figure 6.3.10 : For the group 15, 16, and 17 hydrides, the boiling points for each class of compounds increase with increasing
molecular mass for elements in periods 3, 4, and 5.
If we use this trend to predict the boiling points for the lightest hydride for each group, we would expect NH3 to boil at about −120
°C, H2O to boil at about −80 °C, and HF to boil at about −110 °C. However, when we measure the boiling points for these
compounds, we find that they are dramatically higher than the trends would predict, as shown in Figure 6.3.11. The stark contrast
between our naïve predictions and reality provides compelling evidence for the strength of hydrogen bonding.
6.3.8 https://chem.libretexts.org/@go/page/112709
Figure 6.3.11 : In comparison to periods 3−5, the binary hydrides of period 2 elements in groups 17, 16 and 15 (F, O and N,
respectively) exhibit anomalously high boiling points due to hydrogen bonding.
Exercise 6.3.2
Ethane (CH3CH3) has a melting point of −183 °C and a boiling point of −89 °C. Predict the melting and boiling points for
methylamine (CH3NH2). Explain your reasoning.
Answer:
6.3.9 https://chem.libretexts.org/@go/page/112709
The melting point and boiling point for methylamine are predicted to be significantly greater than those of ethane. CH3CH3
and CH3NH2 are similar in size and mass, but methylamine possesses an −NH group and therefore may exhibit hydrogen
bonding. This greatly increases its IMFs, and therefore its melting and boiling points. It is difficult to predict values, but the
known values are a melting point of −93 °C and a boiling point of −6 °C.
Figure 6.3.12 : Two separate DNA molecules form a double-stranded helix in which the molecules are held together via hydrogen
bonding. (credit: modification of work by Jerome Walker, Dennis Myts)
Each nucleotide contains a (deoxyribose) sugar bound to a phosphate group on one side, and one of four nitrogenous bases on the
other. Two of the bases, cytosine (C) and thymine (T), are single-ringed structures known as pyrimidines. The other two, adenine
(A) and guanine (G), are double-ringed structures called purines. These bases form complementary base pairs consisting of one
purine and one pyrimidine, with adenine pairing with thymine, and cytosine with guanine. Each base pair is held together by
hydrogen bonding. A and T share two hydrogen bonds, C and G share three, and both pairings have a similar shape and structure
Figure 6.3.13
6.3.10 https://chem.libretexts.org/@go/page/112709
Figure 6.3.13 : The geometries of the base molecules result in maximum hydrogen bonding between adenine and thymine (AT)
and between guanine and cytosine (GC), so-called “complementary base pairs.”
The cumulative effect of millions of hydrogen bonds effectively holds the two strands of DNA together. Importantly, the two
strands of DNA can relatively easily “unzip” down the middle since hydrogen bonds are relatively weak compared to the covalent
bonds that hold the atoms of the individual DNA molecules together. This allows both strands to function as a template for
replication.
Summary
Glossary
dipole-dipole attraction
intermolecular attraction between two permanent dipoles
dispersion force
(also, London dispersion force) attraction between two rapidly fluctuating, temporary dipoles; significant only when particles
are very close together
hydrogen bonding
occurs when exceptionally strong dipoles attract; bonding that exists when hydrogen is bonded to one of the three most
electronegative elements: F, O, or N
induced dipole
temporary dipole formed when the electrons of an atom or molecule are distorted by the instantaneous dipole of a neighboring
atom or molecule
instantaneous dipole
temporary dipole that occurs for a brief moment in time when the electrons of an atom or molecule are distributed
asymmetrically
6.3.11 https://chem.libretexts.org/@go/page/112709
intermolecular force
noncovalent attractive force between atoms, molecules, and/or ions
polarizability
measure of the ability of a charge to distort a molecule’s charge distribution (electron cloud)
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
6.3: Intermolecular Forces is shared under a CC BY license and was authored, remixed, and/or curated by LibreTexts.
6.3.12 https://chem.libretexts.org/@go/page/112709
6.3: Intermolecular Forces (Problems)
PROBLEM 6.3.1
In terms of their bulk properties, how do liquids and solids differ? How are they similar?
Answer
Liquids and solids are similar in that they are matter composed of atoms, ions, or molecules. They are incompressible and
have similar densities that are both much larger than those of gases. They are different in that liquids have no fixed shape,
and solids are rigid.
PROBLEM 6.3.2
In terms of the kinetic molecular theory, in what ways are liquids similar to solids? In what ways are liquids different from
solids?
Answer
Particles in liquids and solids are close together, but in liquids they have no regular arrangement or fixed positions.
PROBLEM 6.3.3
In terms of the kinetic molecular theory, in what ways are liquids similar to gases? In what ways are liquids different from
gases?
Answer
They are similar in that the atoms or molecules are free to move from one position to another. They differ in that the particles
of a liquid are confined to the shape of the vessel in which they are placed. In contrast, a gas will expand without limit to fill
the space into which it is placed.
PROBLEM 6.3.4
Explain why liquids assume the shape of any container into which they are poured, whereas solids are rigid and retain their
shape.
Answer
Particles in a solid vibrate about fixed positions and do not generally move in relation to one another; in a liquid, they move
past each other but remain in essentially constant contact
PROBLEM 6.3.5
What is the evidence that all neutral atoms and molecules exert attractive forces on each other?
Answer
All atoms and molecules will condense into a liquid or solid in which the attractive forces exceed the kinetic energy of the
molecules, at sufficiently low temperature.
PROBLEM 6.3.6
The types of intermolecular forces in a substance are identical whether it is a solid, a liquid, or a gas. Why then does a substance
change phase from a gas to a liquid or to a solid?
Answer
Intermolecular forces serve to hold particles close together, whereas the particles’ kinetic energy provides the energy
required to overcome the attractive forces and thus increase the distance between particles. Changes in physical state may be
6.3.1 https://chem.libretexts.org/@go/page/112711
induced by changing the temperature, hence, the average KE, of a given substance
PROBLEM 6.3.7
Why do the boiling points of the noble gases increase in the order He < Ne < Ar < Kr < Xe?
Answer
The London forces typically increase as the number of electrons increase.
PROBLEM 6.3.8
Neon and HF have approximately the same molecular masses.
a. Explain why the boiling points of Neon and HF differ.
b. Compare the change in the boiling points of Ne, Ar, Kr, and Xe with the change of the boiling points of HF, HCl, HBr, and
HI, and explain the difference between the changes with increasing atomic or molecular mass.
Answer a
Ne has only dispersion forces, whereas HF is polar covalent and has hydrogen bonding, dipole-dipole, and dispersion forces.
Answer b
Ne -246 HF 19.5
Xe -108.1 HI -34
The boiling point of the noble gases increases as you increase the molecular weight because of the increasing strength of the
dispersion forces.
For the hydrogen halides, HF does not follow this pattern because it has hydrogen bonding while the other three only has
dipole-dipole interactions.
PROBLEM 6.3.9
Arrange each of the following sets of compounds in order of increasing boiling point temperature:
a. HCl, H2O, SiH4
b. F2, Cl2, Br2
c. CH4, C2H6, C3H8
d. O2, NO, N2
Answer a
SiH4 < HCl < H2O
Answer b
F2 < Cl2 < Br2
Answer c
CH4 < C2H6 < C3H8
Answer d
N2 < O2 < NO
6.3.2 https://chem.libretexts.org/@go/page/112711
PROBLEM 6.3.10
The molecular mass of butanol, C4H9OH, is 74.14; that of ethylene glycol, CH2(OH)CH2OH, is 62.08, yet their boiling points
are 117.2 °C and 174 °C, respectively. Explain the reason for the difference.
Answer
ethylene glycol contains two OH groups which increase the polarity
PROBLEM 6.3.11
On the basis of intermolecular attractions, explain the differences in the boiling points of n–butane (−1 °C) and chloroethane (12
°C), which have similar molar masses.
Answer
Only rather small dipole-dipole interactions from C-H bonds are available to hold n-butane in the liquid state. Chloroethane,
however, has rather large dipole interactions because of the Cl-C bond; the interaction is therefore stronger, leading to a
higher boiling point.
PROBLEM 6.3.12
On the basis of dipole moments and/or hydrogen bonding, explain in a qualitative way the differences in the boiling points of
acetone (56.2 °C) and 1-propanol (97.4 °C), which have similar molar masses.
Answer
1-propanol contains an OH group, which makes it more polar.
PROBLEM 6.3.13
The melting point of H2O(s) is 0 °C. Would you expect the melting point of H2S(s) to be −85 °C, 0 °C, or 185 °C? Explain your
answer.
Answer
−85 °C. Water has stronger hydrogen bonds so it melts at a higher temperature.
PROBLEM 6.3.14
Identify the intermolecular forces present in the following solids:
a. CH3CH2OH
b. CH3CH2CH3
c. CH3CH2Cl
Answer a
hydrogen bonding and dispersion forces
Answer b
dispersion forces
Answer c
dipole-dipole attraction and dispersion forces
6.3.3 https://chem.libretexts.org/@go/page/112711
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Feedback
Think one of the answers above is wrong? Let us know here.
6.3: Intermolecular Forces (Problems) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
6.3.4 https://chem.libretexts.org/@go/page/112711
CHAPTER OVERVIEW
Unit 7: Intermolecular and Intramolecular Forces in Action
Unit Objectives
Distinguish between adhesive and cohesive forces
Define viscosity, surface tension, and capillary rise
Describe the roles of intermolecular attractive forces in each of these properties/phenomena
Define phase transitions and phase transition temperatures
Explain the relation between phase transition temperatures and intermolecular attractive forces
Describe the energetics of covalent and ionic bond formation and breakage
Use the Born-Haber cycle to compute lattice energies for ionic compounds
Use average covalent bond energies to estimate enthalpies of reaction
Unit 7: Intermolecular and Intramolecular Forces in Action is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by LibreTexts.
1
7.1: Surface Tension, Viscosity, and Capillary Action
Skills to Develop
Distinguish between adhesive and cohesive forces
Define viscosity, surface tension, and capillary rise
Describe the roles of intermolecular attractive forces in each of these properties/phenomena
When you pour a glass of water, or fill a car with gasoline, you observe that water and gasoline flow freely. But when you pour
syrup on pancakes or add oil to a car engine, you note that syrup and motor oil do not flow as readily. The viscosity of a liquid is a
measure of its resistance to flow. Water, gasoline, and other liquids that flow freely have a low viscosity. Honey, syrup, motor oil,
and other liquids that do not flow freely, like those shown in Figure 7.1.1, have higher viscosities. We can measure viscosity by
measuring the rate at which a metal ball falls through a liquid (the ball falls more slowly through a more viscous liquid) or by
measuring the rate at which a liquid flows through a narrow tube (more viscous liquids flow more slowly).
Figure 7.1.1 : (a) Honey and (b) motor oil are examples of liquids with high viscosities; they flow slowly. (credit a: modification
of work by Scott Bauer; credit b: modification of work by David Nagy)
The IMFs between the molecules of a liquid, the size and shape of the molecules, and the temperature determine how easily a
liquid flows. As Table 7.1.1 shows, the more structurally complex are the molecules in a liquid and the stronger the IMFs between
them, the more difficult it is for them to move past each other and the greater is the viscosity of the liquid. As the temperature
increases, the molecules move more rapidly and their kinetic energies are better able to overcome the forces that hold them
together; thus, the viscosity of the liquid decreases.
Table 7.1.1: Viscosities of Common Substances at 25 °C
Substance Formula Viscosity (mPa·s)
mercury Hg 1.526
7.1.1 https://chem.libretexts.org/@go/page/119785
The various IMFs between identical molecules of a substance are examples of cohesive forces. The molecules within a liquid are
surrounded by other molecules and are attracted equally in all directions by the cohesive forces within the liquid. However, the
molecules on the surface of a liquid are attracted only by about one-half as many molecules. Because of the unbalanced molecular
attractions on the surface molecules, liquids contract to form a shape that minimizes the number of molecules on the surface—that
is, the shape with the minimum surface area. A small drop of liquid tends to assume a spherical shape, as shown in Figure 7.1.2,
because in a sphere, the ratio of surface area to volume is at a minimum. Larger drops are more greatly affected by gravity, air
resistance, surface interactions, and so on, and as a result, are less spherical.
Figure 7.1.2 : Attractive forces result in a spherical water drop that minimizes surface area; cohesive forces hold the sphere
together; adhesive forces keep the drop attached to the web. (credit photo: modification of work by “OliBac”/Flickr)
Surface tension is defined as the energy required to increase the surface area of a liquid, or the force required to increase the length
of a liquid surface by a given amount. This property results from the cohesive forces between molecules at the surface of a liquid,
and it causes the surface of a liquid to behave like a stretched rubber membrane. Surface tensions of several liquids are presented in
Table 7.1.2.
Table 7.1.2: Surface Tensions of Common Substances at 25 °C
Substance Formula Surface Tension (mN/m)
mercury Hg 458.48
Among common liquids, water exhibits a distinctly high surface tension due to strong hydrogen bonding between its molecules. As
a result of this high surface tension, the surface of water represents a relatively “tough skin” that can withstand considerable force
without breaking. A steel needle carefully placed on water will float. Some insects, like the one shown in Figure 7.1.3, even though
they are denser than water, move on its surface because they are supported by the surface tension.
Figure 7.1.3 : Surface tension (right) prevents this insect, a “water strider,” from sinking into the water.
The IMFs of attraction between two different molecules are called adhesive forces. Consider what happens when water comes into
contact with some surface. If the adhesive forces between water molecules and the molecules of the surface are weak compared to
the cohesive forces between the water molecules, the water does not “wet” the surface. For example, water does not wet waxed
7.1.2 https://chem.libretexts.org/@go/page/119785
surfaces or many plastics such as polyethylene. Water forms drops on these surfaces because the cohesive forces within the drops
are greater than the adhesive forces between the water and the plastic. Water spreads out on glass because the adhesive force
between water and glass is greater than the cohesive forces within the water. When water is confined in a glass tube, its meniscus
(surface) has a concave shape because the water wets the glass and creeps up the side of the tube. On the other hand, the cohesive
forces between mercury atoms are much greater than the adhesive forces between mercury and glass. Mercury therefore does not
wet glass, and it forms a convex meniscus when confined in a tube because the cohesive forces within the mercury tend to draw it
into a drop (Figure 7.1.4).
Figure 7.1.4 : Differences in the relative strengths of cohesive and adhesive forces result in different meniscus shapes for mercury
(left) and water (right) in glass tubes. (credit: Mark Ott)
If you place one end of a paper towel in spilled wine, as shown in Figure 7.1.5, the liquid wicks up the paper towel. A similar
process occurs in a cloth towel when you use it to dry off after a shower. These are examples of capillary action—when a liquid
flows within a porous material due to the attraction of the liquid molecules to the surface of the material and to other liquid
molecules. The adhesive forces between the liquid and the porous material, combined with the cohesive forces within the liquid,
may be strong enough to move the liquid upward against gravity.
Figure 7.1.5 : Wine wicks up a paper towel (left) because of the strong attractions of water (and ethanol) molecules to the −OH
groups on the towel’s cellulose fibers and the strong attractions of water molecules to other water (and ethanol) molecules (right).
(credit photo: modification of work by Mark Blaser)
Towels soak up liquids like water because the fibers of a towel are made of molecules that are attracted to water molecules. Most
cloth towels are made of cotton, and paper towels are generally made from paper pulp. Both consist of long molecules of cellulose
that contain many −OH groups. Water molecules are attracted to these −OH groups and form hydrogen bonds with them, which
draws the H2O molecules up the cellulose molecules. The water molecules are also attracted to each other, so large amounts of
water are drawn up the cellulose fibers.
Capillary action can also occur when one end of a small diameter tube is immersed in a liquid, as illustrated in Figure 7.1.6. If the
liquid molecules are strongly attracted to the tube molecules, the liquid creeps up the inside of the tube until the weight of the liquid
and the adhesive forces are in balance. The smaller the diameter of the tube is, the higher the liquid climbs. It is partly by capillary
action occurring in plant cells called xylem that water and dissolved nutrients are brought from the soil up through the roots and
into a plant. Capillary action is the basis for thin layer chromatography, a laboratory technique commonly used to separate small
quantities of mixtures. You depend on a constant supply of tears to keep your eyes lubricated and on capillary action to pump tear
fluid away.
7.1.3 https://chem.libretexts.org/@go/page/119785
Figure 7.1.6 : Depending upon the relative strengths of adhesive and cohesive forces, a liquid may rise (such as water) or fall
(such as mercury) in a glass capillary tube. The extent of the rise (or fall) is directly proportional to the surface tension of the
liquid and inversely proportional to the density of the liquid and the radius of the tube.
The height to which a liquid will rise in a capillary tube is determined by several factors as shown in the following equation:
2T cos θ
h = (7.1.1)
rρg
where
h is the height of the liquid inside the capillary tube relative to the surface of the liquid outside the tube,
T is the surface tension of the liquid,
θ is the contact angle between the liquid and the tube,
r is the radius of the tube, ρ is the density of the liquid, and
g is the acceleration due to gravity, 9.8 m/s2.
When the tube is made of a material to which the liquid molecules are strongly attracted, they will spread out completely on the
surface, which corresponds to a contact angle of 0°. This is the situation for water rising in a glass tube. We will not concern
ourselves with calculating capillary height in this course.
Applications: Capillary Action is Used to Draw Blood
Many medical tests require drawing a small amount of blood, for example to determine the amount of glucose in someone with
diabetes or the hematocrit level in an athlete. This procedure can be easily done because of capillary action, the ability of a
liquid to flow up a small tube against gravity, as shown in Figure 7.1.7. When your finger is pricked, a drop of blood forms and
holds together due to surface tension—the unbalanced intermolecular attractions at the surface of the drop. Then, when the open
end of a narrow-diameter glass tube touches the drop of blood, the adhesive forces between the molecules in the blood and those
at the glass surface draw the blood up the tube. How far the blood goes up the tube depends on the diameter of the tube (and the
type of fluid). A small tube has a relatively large surface area for a given volume of blood, which results in larger (relative)
attractive forces, allowing the blood to be drawn farther up the tube. The liquid itself is held together by its own cohesive forces.
When the weight of the liquid in the tube generates a downward force equal to the upward force associated with capillary action,
the liquid stops rising.
Figure 7.1.7 :: Blood is collected for medical analysis by capillary action, which draws blood into a small diameter glass tube.
(credit: modification of work by Centers for Disease Control and Prevention)
7.1.4 https://chem.libretexts.org/@go/page/119785
Summary
Video 7.1.1 : An overview of intermolecular forces in action as surface tension, viscosity, and capillary action.
The intermolecular forces between molecules in the liquid state vary depending upon their chemical identities and result in
corresponding variations in various physical properties. Cohesive forces between like molecules are responsible for a liquid’s
viscosity (resistance to flow) and surface tension (elasticity of a liquid surface). Adhesive forces between the molecules of a liquid
and different molecules composing a surface in contact with the liquid are responsible for phenomena such as surface wetting and
capillary rise.
Learn More: Why does ice float in water?
Video 7.1.2 : Ice floating in water is one of the unique properties of water.
Glossary
adhesive force
force of attraction between molecules of different chemical identities
capillary action
flow of liquid within a porous material due to the attraction of the liquid molecules to the surface of the material and to other
liquid molecules
cohesive force
force of attraction between identical molecules
surface tension
energy required to increase the area, or length, of a liquid surface by a given amount
7.1.5 https://chem.libretexts.org/@go/page/119785
viscosity
measure of a liquid’s resistance to flow
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
TED-Ed’s commitment to creating lessons worth sharing is an extension of TED’s mission of spreading great ideas. Within
TED-Ed’s growing library of TED-Ed animations, you will find carefully curated educational videos, many of which represent
collaborations between talented educators and animators nominated through the TED-Ed website.
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
7.1: Surface Tension, Viscosity, and Capillary Action is shared under a CC BY license and was authored, remixed, and/or curated by LibreTexts.
7.1.6 https://chem.libretexts.org/@go/page/119785
7.1: Surface Tension, Viscosity, and Capillary Action (Problems)
PROBLEM 7.1.1
The test tubes shown here contain equal amounts of the specified motor oils. Identical metal spheres were dropped at the same
time into each of the tubes, and a brief moment later, the spheres had fallen to the heights indicated in the illustration. Rank the
motor oils in order of increasing viscosity, and explain your reasoning:
Answer
20 < 30 < 40 < 50
PROBLEM 7.1.2
Although steel is denser than water, a steel needle or paper clip placed carefully lengthwise on the surface of still water can be
made to float. Explain at a molecular level how this is possible:
Answer
The water molecules have strong intermolecular forces of hydrogen bonding. The water molecules are thus attracted strongly
to one another and exhibit a relatively large surface tension, forming a type of “skin” at its surface. This skin can support a
bug or paper clip if gently placed on the water.
PROBLEM 7.1.3
The surface tension and viscosity values for diethyl ether, acetone, ethanol, and ethylene glycol are shown here.
7.1.1 https://chem.libretexts.org/@go/page/119786
a. Explain their differences in viscosity in terms of the size and shape of their molecules and their IMFs.
b. Explain their differences in surface tension in terms of the size and shape of their molecules and their IMFs.
Answer a
The viscosity increases as the molecular weight (size) of the molecules increases. Additionally, the more polar the molecule,
the more viscous.
Answer b
The surface tension increases as the molecular weight of the molecule increases. Additionally, the more polar the molecule,
the higher the surface tension.
PROBLEM 7.1.4
You may have heard someone use the figure of speech “slower than molasses in winter” to describe a process that occurs slowly.
Explain why this is an apt idiom, using concepts of molecular size and shape, molecular interactions, and the effect of changing
temperature.
Answer
Temperature has an effect on intermolecular forces: the higher the temperature, the greater the kinetic energies of the
molecules and the greater the extent to which their intermolecular forces are overcome, and so the more fluid (less viscous)
the liquid; the lower the temperature, the lesser the intermolecular forces are overcome, and so the more viscous the liquid.
PROBLEM 7.1.5
It is often recommended that you let your car engine run idle to warm up before driving, especially on cold winter days. While
the benefit of prolonged idling is dubious, it is certainly true that a warm engine is more fuel efficient than a cold one. Explain
the reason for this.
Answer
The fluids in the engine are warmer, decreasing their viscosity, which aids in lubricating the moving parts of the engine
causing it to run smoother.
PROBLEM 7.1.6
The surface tension and viscosity of water at several different temperatures are given in this table.
0 °C 75.6 1.79
20 °C 72.8 1.00
60 °C 66.2 0.47
7.1.2 https://chem.libretexts.org/@go/page/119786
Water Surface Tension (mN/m) Viscosity (mPa s)
a. As temperature increases, what happens to the surface tension of water? Explain why this occurs, in terms of molecular
interactions and the effect of changing temperature.
b. As temperature increases, what happens to the viscosity of water? Explain why this occurs, in terms of molecular
interactions and the effect of changing temperature.
Answer a
As the water reaches higher temperatures, the increased kinetic energies of its molecules are more effective in overcoming
hydrogen bonding, and so its surface tension decreases. Surface tension and intermolecular forces are directly related.
Answer b
The same trend in viscosity is seen as in surface tension, and for the same reason.
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Feedback
Think one of the answers above is wrong? Let us know here.
7.1: Surface Tension, Viscosity, and Capillary Action (Problems) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by LibreTexts.
7.1.3 https://chem.libretexts.org/@go/page/119786
7.2: Vapor Pressure
Skills to Develop
Define phase transitions and phase transition temperatures
Explain the relation between phase transition temperatures and intermolecular attractive forces
We witness and utilize changes of physical state, or phase transitions, in a great number of ways. As one example of global
significance, consider the evaporation, condensation, freezing, and melting of water. These changes of state are essential aspects of
our earth’s water cycle as well as many other natural phenomena and technological processes of central importance to our lives. In
this module, the essential aspects of phase transitions are explored.
Figure 7.2.1 : In a closed container, dynamic equilibrium is reached when (a) the rate of molecules escaping from the liquid to
become the gas (b) increases and eventually (c) equals the rate of gas molecules entering the liquid. When this equilibrium is
reached, the vapor pressure of the gas is constant, although the vaporization and condensation processes continue.
The chemical identities of the molecules in a liquid determine the types (and strengths) of intermolecular attractions possible;
consequently, different substances will exhibit different equilibrium vapor pressures. Relatively strong intermolecular attractive
forces will serve to impede vaporization as well as favoring “recapture” of gas-phase molecules when they collide with the liquid
surface, resulting in a relatively low vapor pressure. Weak intermolecular attractions present less of a barrier to vaporization, and a
reduced likelihood of gas recapture, yielding relatively high vapor pressures. The following example illustrates this dependence of
vapor pressure on intermolecular attractive forces.
Solution
7.2.1 https://chem.libretexts.org/@go/page/119787
Diethyl ether has a very small dipole and most of its intermolecular attractions are London forces. Although this molecule is the
largest of the four under consideration, its IMFs are the weakest and, as a result, its molecules most readily escape from the
liquid. It also has the highest vapor pressure. Due to its smaller size, ethanol exhibits weaker dispersion forces than diethyl ether.
However, ethanol is capable of hydrogen bonding and, therefore, exhibits stronger overall IMFs, which means that fewer
molecules escape from the liquid at any given temperature, and so ethanol has a lower vapor pressure than diethyl ether. Water
is much smaller than either of the previous substances and exhibits weaker dispersion forces, but its extensive hydrogen bonding
provides stronger intermolecular attractions, fewer molecules escaping the liquid, and a lower vapor pressure than for either
diethyl ether or ethanol. Ethylene glycol has two −OH groups, so, like water, it exhibits extensive hydrogen bonding. It is much
larger than water and thus experiences larger London forces. Its overall IMFs are the largest of these four substances, which
means its vaporization rate will be the slowest and, consequently, its vapor pressure the lowest.
Exercise 7.2.1
At 20 °C, the vapor pressures of several alcohols are given in this table. Explain these vapor pressures in terms of types and
extents of IMFs for these alcohols:
Vapor Pressure at 20 °C 11.9 kPa 5.95 kPa 2.67 kPa 0.56 kPa
Answer
All these compounds exhibit hydrogen bonding; these strong IMFs are difficult for the molecules to overcome, so the vapor
pressures are relatively low. As the size of molecule increases from methanol to butanol, dispersion forces increase, which
means that the vapor pressures decrease as observed:
As temperature increases, the vapor pressure of a liquid also increases due to the increased average KE of its molecules. Recall that
at any given temperature, the molecules of a substance experience a range of kinetic energies, with a certain fraction of molecules
having a sufficient energy to overcome IMF and escape the liquid (vaporize). At a higher temperature, a greater fraction of
molecules have enough energy to escape from the liquid, as shown in Figure 7.2.2. The escape of more molecules per unit of time
and the greater average speed of the molecules that escape both contribute to the higher vapor pressure.
Figure 7.2.2 : Temperature affects the distribution of kinetic energies for the molecules in a liquid. At the higher temperature,
more molecules have the necessary kinetic energy, KE, to escape from the liquid into the gas phase.
Boiling Points
7.2.2 https://chem.libretexts.org/@go/page/119787
BOILING the Liquid Metal - MERCURY (…
(…
Figure 7.2.3 : The boiling points of liquids are the temperatures at which their equilibrium vapor pressures equal the pressure of
the surrounding atmosphere. Normal boiling points are those corresponding to a pressure of 1 atm (101.3 kPa.)
7.2.3 https://chem.libretexts.org/@go/page/119787
Solution
The graph of the vapor pressure of water versus temperature in Figure 7.2.3 indicates that the vapor pressure of water is 68 kPa
at about 90 °C. Thus, at about 90 °C, the vapor pressure of water will equal the atmospheric pressure in Leadville, and water
will boil.
Exercise 7.2.2
The boiling point of ethyl ether was measured to be 10 °C at a base camp on the slopes of Mount Everest. Use Figure 7.2.3 to
determine the approximate atmospheric pressure at the camp.
Answer
Approximately 40 kPa (0.4 atm)
The quantitative relation between a substance’s vapor pressure and its temperature is described by the Clausius-Clapeyron
equation:
−Δ Hvap /RT
P = Ae (7.2.1)
where
ΔHvap is the enthalpy of vaporization for the liquid,
R is the gas constant, and
ln A is a constant whose value depends on the chemical identity of the substance.
Equation 7.2.1 is often rearranged into logarithmic form to yield the linear equation:
ΔHvap
ln P = − + ln A (7.2.2)
RT
This linear equation may be expressed in a two-point format that is convenient for use in various computations, as demonstrated in
the examples and exercises that follow. If at temperature T , the vapor pressure is P , and at temperature T , the vapor pressure is
1 1 2
ΔHvap
ln P1 = − + ln A (7.2.3)
RT1
and
ΔHvap
ln P2 = − + ln A (7.2.4)
RT2
Since the constant, ln A, is the same, these two equations may be rearranged to isolate ln A and then set them equal to one another:
ΔHvap ΔHvap
ln P1 + = ln P2 +
RT1 RT2
7.2.4 https://chem.libretexts.org/@go/page/119787
P2 ΔHvap 1 1
ln( ) = ( − )
P1 R T1 T2
P1 = 10.0 kP a and
o
T2 = 98.8 C = 372.0 K
P2 = 100 kP a
we can substitute them into this equation and solve for ΔHvap . Rearranging the Clausius-Clapeyron equation and solving for
ΔH vapyields:
P2
R ⋅ ln( )
P1
ΔHvap =
1 1
( − )
T1 T2
100 kP a
(8.3145 J/mol ⋅ K) ⋅ ln( )
10.0 kP a
=
1 1
( − )
307.2 K 372.0 K
Note that the pressure can be in any units, so long as they agree for both P values, but the temperature must be in kelvin for the
Clausius-Clapeyron equation to be valid.
Exercise 7.2.3
At 20.0 °C, the vapor pressure of ethanol is 5.95 kPa, and at 63.5 °C, its vapor pressure is 53.3 kPa. Use this information to
estimate the enthalpy of vaporization for ethanol.
Answer
47,782 J/mol = 47.8 kJ/mol
P2 ΔHvap 1 1
ln( ) = ( − )
P1 R T1 T2
Since the normal boiling point is the temperature at which the vapor pressure equals atmospheric pressure at sea level, we know
one vapor pressure-temperature value (T = 80.1 °C = 353.3 K, P = 101.3 kPa, ΔH
1 1 = 30.8 kJ/mol) and want to find the
vap
temperature (T ) that corresponds to vapor pressure P2 = 83.4 kPa. We can substitute these values into the Clausius-Clapeyron
2
equation and then solve for T . Rearranging the Clausius-Clapeyron equation and solving for T yields:
2 2
7.2.5 https://chem.libretexts.org/@go/page/119787
−1
P2
⎛ ⎞
−R ⋅ ln( )
⎜ P1 1 ⎟
T2 = ⎜ + ⎟
⎜ ΔHvap T1 ⎟
⎝ ⎠
−1
83.4 kPa
⎛ ⎞
−(8.3145 J/mol ⋅ K) ⋅ ln( )
⎜ 101.3 kPa 1 ⎟
= ⎜ + ⎟
⎜ 30, 800 J/mol 353.3 K ⎟
⎝ ⎠
∘
= 346.9 K or 73.8 C
Exercise 7.2.4
For acetone (CH ) CO, the normal boiling point is 56.5 °C and the enthalpy of vaporization is 31.3 kJ/mol. What is the vapor
3 2
Answer
30.1 kPa
Enthalpy of Vaporization
Vaporization is an endothermic process. The cooling effect can be evident when you leave a swimming pool or a shower. When the
water on your skin evaporates, it removes heat from your skin and causes you to feel cold. The energy change associated with the
vaporization process is the enthalpy of vaporization, ΔH . For example, the vaporization of water at standard temperature is
vap
represented by:
As described in the chapter on thermochemistry, the reverse of an endothermic process is exothermic. And so, the condensation of
a gas releases heat:
7.2.6 https://chem.libretexts.org/@go/page/119787
Figure 7.2.4 : Evaporation of sweat helps cool the body. (credit: “Kullez”/Flickr)
Solution We start with the known volume of sweat (approximated as just water) and use the given information to convert to the
amount of heat needed:
1000 g 1 mol 43.46 kJ
3
1.5 L × × × = 3.6 × 10 kJ
1 L 18 g 1 mol
Answer
28 kJ
Figure 7.2.5 : (a) This beaker of ice has a temperature of −12.0 °C. (b) After 10 minutes the ice has absorbed enough heat from
the air to warm to 0 °C. A small amount has melted. (c) Thirty minutes later, the ice has absorbed more heat, but its temperature is
still 0 °C. The ice melts without changing its temperature. (d) Only after all the ice has melted does the heat absorbed cause the
temperature to increase to 22.2 °C. (credit: modification of work by Mark Ott).
If we stop heating during melting and place the mixture of solid and liquid in a perfectly insulated container so no heat can enter or
escape, the solid and liquid phases remain in equilibrium. This is almost the situation with a mixture of ice and water in a very good
thermos bottle; almost no heat gets in or out, and the mixture of solid ice and liquid water remains for hours. In a mixture of solid
and liquid at equilibrium, the reciprocal processes of melting and freezing occur at equal rates, and the quantities of solid and liquid
7.2.7 https://chem.libretexts.org/@go/page/119787
therefore remain constant. The temperature at which the solid and liquid phases of a given substance are in equilibrium is called the
melting point of the solid or the freezing point of the liquid. Use of one term or the other is normally dictated by the direction of the
phase transition being considered, for example, solid to liquid (melting) or liquid to solid (freezing).
The enthalpy of fusion and the melting point of a crystalline solid depend on the strength of the attractive forces between the units
present in the crystal. Molecules with weak attractive forces form crystals with low melting points. Crystals consisting of particles
with stronger attractive forces melt at higher temperatures.
The amount of heat required to change one mole of a substance from the solid state to the liquid state is the enthalpy of fusion,
ΔHfus of the substance. The enthalpy of fusion of ice is 6.0 kJ/mol at 0 °C. Fusion (melting) is an endothermic process:
H O(s) → H O(l) ΔHfus = 6.01 kJ/mol (7.2.8)
2 2
The reciprocal process, freezing, is an exothermic process whose enthalpy change is −6.0 kJ/mol at 0 °C:
H O(l) → H O(s) ΔHfrz = −ΔHfus = −6.01 kJ/mol (7.2.9)
2 2
Figure 7.2.6 : Sublimation of solid iodine in the bottom of the tube produces a purple gas that subsequently deposits as solid
iodine on the colder part of the tube above. (credit: modification of work by Mark Ott)
Like vaporization, the process of sublimation requires an input of energy to overcome intermolecular attractions. The enthalpy of
sublimation, ΔHsub, is the energy required to convert one mole of a substance from the solid to the gaseous state. For example, the
sublimation of carbon dioxide is represented by:
Likewise, the enthalpy change for the reverse process of deposition is equal in magnitude but opposite in sign to that for
sublimation:
CO (g) ⟶ CO (s) ΔHdep = −ΔHsub = −26.1 kJ/mol (7.2.11)
2 2
Consider the extent to which intermolecular attractions must be overcome to achieve a given phase transition. Converting a solid
into a liquid requires that these attractions be only partially overcome; transition to the gaseous state requires that they be
completely overcome. As a result, the enthalpy of fusion for a substance is less than its enthalpy of vaporization. This same logic
can be used to derive an approximate relation between the enthalpies of all phase changes for a given substance. Though not an
entirely accurate description, sublimation may be conveniently modeled as a sequential two-step process of melting followed by
vaporization in order to apply Hess’s Law.
7.2.8 https://chem.libretexts.org/@go/page/119787
solid ⟶ liquid ΔHfus (7.2.12)
Viewed in this manner, the enthalpy of sublimation for a substance may be estimated as the sum of its enthalpies of fusion and
vaporization, as illustrated in Figure 7.2.7. For example:
Figure 7.2.7 : For a given substance, the sum of its enthalpy of fusion and enthalpy of vaporization is approximately equal to its
enthalpy of sublimation.
Summary
Key Equations
−Δ Hv ap /RT
P = Ae
ΔHvap
ln P = − + ln A
RT
P2 ΔHvap 1 1
ln( ) = ( − )
P1 R T1 T2
7.2.9 https://chem.libretexts.org/@go/page/119787
Glossary
boiling point
temperature at which the vapor pressure of a liquid equals the pressure of the gas above it
Clausius-Clapeyron equation
mathematical relationship between the temperature, vapor pressure, and enthalpy of vaporization for a substance
condensation
change from a gaseous to a liquid state
deposition
change from a gaseous state directly to a solid state
dynamic equilibrium
state of a system in which reciprocal processes are occurring at equal rates
freezing
change from a liquid state to a solid state
freezing point
temperature at which the solid and liquid phases of a substance are in equilibrium; see also melting point
melting
change from a solid state to a liquid state
melting point
temperature at which the solid and liquid phases of a substance are in equilibrium; see also freezing point
sublimation
change from solid state directly to gaseous state
vapor pressure
(also, equilibrium vapor pressure) pressure exerted by a vapor in equilibrium with a solid or a liquid at a given temperature
vaporization
change from liquid state to gaseous state
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
Guillotined Chemistry is a production of Mark Anticole and available for free on Youtube.
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
7.2.10 https://chem.libretexts.org/@go/page/119787
7.2: Vapor Pressure is shared under a CC BY license and was authored, remixed, and/or curated by LibreTexts.
7.2.11 https://chem.libretexts.org/@go/page/119787
7.2: Vapor Pressure (Problems)
PROBLEM 7.2.1
Heat is added to boiling water. Explain why the temperature of the boiling water does not change. What does change?
Answer
The heat is absorbed by the water, providing the energy required to partially overcome intermolecular attractive forces in the
liquid and causing a phase transition to gaseous water. The solution remains at 100 °C until all the water is boiled. Only the
amount of water existing as liquid water changes until the liquid disappears. Then the temperature of the water can rise.
PROBLEM 7.2.2
Heat is added to ice at 0 °C. Explain why the temperature of the ice does not change. What does change?
Answer
The heat is absorbed by the ice, providing the energy required to partially overcome intermolecular attractive forces in the
solid and causing a phase transition to liquid water. The solution remains at 0 °C until all the ice is melted. Only the amount
of water existing as ice changes until the ice disappears. Then the temperature of the water can rise.
PROBLEM 7.2.3
What is the relationship between the intermolecular forces in a liquid and its vapor pressure?
Answer
The vapor pressure of a liquid decreases as the strength of its intermolecular forces increases.
PROBLEM 7.2.4
What is the relationship between the intermolecular forces in a solid and its melting temperature?
Answer
The stronger the intermolecular forces, the higher the melting point.
PROBLEM 7.2.5
Why does spilled gasoline evaporate more rapidly on a hot day than on a cold day?
Answer
As the temperature increases, the average kinetic energy of the molecules of gasoline increases and so a greater fraction of
molecules have sufficient energy to escape from the liquid than at lower temperatures.
PROBLEM 7.2.6
When is the boiling point of a liquid equal to its normal boiling point?
Answer
When the pressure of gas above the liquid is exactly 1 atm
PROBLEM 7.2.7
How does the boiling of a liquid differ from its evaporation?
Answer
boiling implies an application of energy in the form of heat.
7.2.1 https://chem.libretexts.org/@go/page/119788
PROBLEM 7.2.8
Explain the following observations:
a. It takes longer to cook an egg in Klamath Falls, Oregon (altitude, 4200 feet above sea level) than it does in Boston (at sea
level).
b. Perspiring is a mechanism for cooling the body.
Answer a
At 4200 feet, the atmospheric pressure is lower than at sea level, and water will therefore boil at a lower temperature. This
lower temperature will cause the physical and chemical changes involved in cooking the egg to proceed more slowly, and a
longer time is required to fully cook the egg.
Answer b
As long as the air surrounding the body contains less water vapor than the maximum that air can hold at that temperature,
perspiration will evaporate, thereby cooling the body by removing the heat of vaporization required to vaporize the water.
PROBLEM 7.2.9
Explain why the molar enthalpies of vaporization of the following substances increase in the order CH4 < C2H6 < C3H8, even
though all three substances experience the same dispersion forces when in the liquid state.
Answer
Dispersion forces increase with molecular mass or size. As the number of atoms composing the molecules in this
homologous series increases, so does the extent of intermolecular attraction via dispersion forces and, consequently, the
energy required to overcome these forces and vaporize the liquids.
PROBLEM 7.2.10
Explain why the enthalpies of vaporization of the following substances increase in the order CH4 < NH3 < H2O, even though all
three substances have approximately the same molar mass.
Answer
CH4 is non-polar (dispersion forces only)
NH3 is polar (dipole-dipole interactions)
H2O is polar (hydrogen bonding)
PROBLEM 7.2.11
Which contains the compounds listed correctly in order of increasing boiling points?
a. N2 < CS2 < H2O < KCl
b. H2O < N2 < CS2 < KCl
c. N2 < KCl < CS2 < H2O
d. CS2 < N2 < KCl < H2O
e. KCl < H2O < CS2 < N2
Answer
a is correct; N2 is nonpolar, CS2 is polar (dipole-dipole), H2O is polar (H-bonding), KCl is ionic (ion-dipole)
7.2.2 https://chem.libretexts.org/@go/page/119788
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Feedback
Think one of the answers above is wrong? Let us know here.
7.2: Vapor Pressure (Problems) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
7.2.3 https://chem.libretexts.org/@go/page/119788
7.3: Ionic Bond Formation and Strength
Skills to Develop
Explain the formation of cations, anions, and ionic compounds
Describe the energetics of ionic bond formation and breakage
Use the Born-Haber cycle to compute lattice energies for ionic compounds
A bond’s strength describes how strongly each atom is joined to another atom, and therefore how much energy is required to break
the bond between the two atoms. In a previous section, you learned about the bond strength of covalent bonds. Now we
will compare that to the strength of ionic bonds, which is related to the lattice energy of a compound. But first, let's explore how
ionic bonds form.
7.3.1 https://chem.libretexts.org/@go/page/119789
dissolve readily in water. Once dissolved or melted, ionic compounds are excellent conductors of electricity and heat because the
ions can move about freely.
Neutral atoms and their associated ions have very different physical and chemical properties. Sodium atoms form sodium metal, a
soft, silvery-white metal that burns vigorously in air and reacts explosively with water. Chlorine atoms form chlorine gas, Cl2, a
yellow-green gas that is extremely corrosive to most metals and very poisonous to animals and plants. The vigorous reaction
between the elements sodium and chlorine forms the white, crystalline compound sodium chloride, common table salt, which
contains sodium cations and chloride anions (Figure Figure 7.3.1). The compound composed of these ions exhibits properties
entirely different from the properties of the elements sodium and chlorine. Chlorine is poisonous, but sodium chloride is essential
to life; sodium atoms react vigorously with water, but sodium chloride simply dissolves in water.
Figure 7.3.1 : (a) Sodium is a soft metal that must be stored in mineral oil to prevent reaction with air or water. (b) Chlorine is a
pale yellow-green gas. (c) When combined, they form white crystals of sodium chloride (table salt). (credit a: modification of work
by “Jurii”/Wikimedia Commons)
Binary ionic compounds are composed of just two elements: a metal (which forms the cations) and a nonmetal (which forms the
anions). For example, NaCl is a binary ionic compound. We can think about the formation of such compounds in terms of the
periodic properties of the elements. Many metallic elements have relatively low ionization potentials and lose electrons easily.
These elements lie to the left in a period or near the bottom of a group on the periodic table. Nonmetal atoms have relatively high
electron affinities and thus readily gain electrons lost by metal atoms, thereby filling their valence shells. Nonmetallic elements are
found in the upper-right corner of the periodic table.
As all substances must be electrically neutral, the total number of positive charges on the cations of an ionic compound must equal
the total number of negative charges on its anions. The formula of an ionic compound represents the simplest ratio of the numbers
of ions necessary to give identical numbers of positive and negative charges. For example, the formula for aluminum oxide, Al2O3,
indicates that this ionic compound contains two aluminum cations, Al3+, for every three oxide anions, O2− [thus, (2 × +3) + (3 × –
2) = 0].
It is important to note, however, that the formula for an ionic compound does not represent the physical arrangement of its ions. It
is incorrect to refer to a sodium chloride (NaCl) “molecule” because there is not a single ionic bond, per se, between any specific
pair of sodium and chloride ions. The attractive forces between ions are isotropic—the same in all directions—meaning that any
particular ion is equally attracted to all of the nearby ions of opposite charge. This results in the ions arranging themselves into a
tightly bound, three-dimensional lattice structure. Sodium chloride, for example, consists of a regular arrangement of equal
numbers of Na+ cations and Cl– anions (Figure Figure 7.3.2).
Two diagrams are shown and labeled “a” and “b.” Diagram a shows a cube made up of twenty-seven alternating purple and green spheres. The purple spheres are smaller than the green spheres. Diagram b shows the
same spheres, but this time, they are spread out and connected in three dimensions by white rods. The purple spheres are labeled “N superscript postive sign” while the green are labeled “C l superscript negative sign.”
Figure 7.3.2 : The atoms in sodium chloride (common table salt) are arranged to (a) maximize opposite charges interacting. The
smaller spheres represent sodium ions, the larger ones represent chloride ions. In the expanded view (b), the geometry can be seen
more clearly. Note that each ion is “bonded” to all of the surrounding ions—six in this case.
The strong electrostatic attraction between Na+ and Cl– ions holds them tightly together in solid NaCl. It requires 769 kJ of energy
to dissociate one mole of solid NaCl into separate gaseous Na+ and Cl– ions:
+ −
NaCl(s) ⟶ Na (g) + Cl (g) ΔH = 769 kJ (7.3.1)
7.3.2 https://chem.libretexts.org/@go/page/119789
What are Ionic Bonds? | Properties of M…
M…
energy required to separate one mole of the solid into its component gaseous ions. For the ionic solid MX, the lattice energy is the
enthalpy change of the process:
+ n−
M X(s) ⟶ M n +X ΔHlattice (7.3.2)
(g) (g)
Note that we are using the convention where the ionic solid is separated into ions, so our lattice energies will be endothermic
(positive values). Some texts use the equivalent but opposite convention, defining lattice energy as the energy released when
separate ions combine to form a lattice and giving negative (exothermic) values. Thus, if you are looking up lattice energies in
another reference, be certain to check which definition is being used. In both cases, a larger magnitude for lattice energy indicates a
more stable ionic compound. For sodium chloride, ΔHlattice = 769 kJ. Thus, it requires 769 kJ to separate one mole of solid NaCl
into gaseous Na+ and Cl– ions. When one mole each of gaseous Na+ and Cl– ions form solid NaCl, 769 kJ of heat is released.
The lattice energy ΔH of an ionic crystal can be expressed by the following equation (derived from Coulomb’s law,
lattice
in which
C is a constant that depends on the type of crystal structure;
Z and Z are the charges on the ions; and
+ –
R is the interionic distance (the sum of the radii of the positive and negative ions).
o
Thus, the lattice energy of an ionic crystal increases rapidly as the charges of the ions increase and the sizes of the ions decrease.
When all other parameters are kept constant, doubling the charge of both the cation and anion quadruples the lattice energy. For
example, the lattice energy of LiF (Z+ and Z– = 1) is 1023 kJ/mol, whereas that of MgO (Z+ and Z– = 2) is 3900 kJ/mol (Ro is
nearly the same—about 200 pm for both compounds).
Different interatomic distances produce different lattice energies. For example, we can compare the lattice energy of MgF2 (2957
kJ/mol) to that of MgI2 (2327 kJ/mol) to observe the effect on lattice energy of the smaller ionic size of F– as compared to I–.
7.3.3 https://chem.libretexts.org/@go/page/119789
In these two ionic compounds, the charges Z+ and Z– are the same, so the difference in lattice energy will mainly depend upon
Ro. The O2– ion is smaller than the Se2– ion. Thus, Al2O3 would have a shorter interionic distance than Al2Se3, and Al2O3 would
have the larger lattice energy.
Exercise 7.3.2
Zinc oxide, ZnO, is a very effective sunscreen. How would the lattice energy of ZnO compare to that of NaCl?
Answer
ZnO would have the larger lattice energy because the Z values of both the cation and the anion in ZnO are greater, and the
interionic distance of ZnO is smaller than that of NaCl.
Figure 7.3.3 : The Born-Haber cycle shows the relative energies of each step involved in the formation of an ionic solid from the
necessary elements in their reference states.
We begin with the elements in their most common states, Cs(s) and F2(g). The ΔH represents the conversion of solid cesium into
∘
s
a gas, and then the ionization energy converts the gaseous cesium atoms into cations. In the next step, we account for the energy
required to break the F–F bond to produce fluorine atoms. Converting one mole of fluorine atoms into fluoride ions is an
exothermic process, so this step gives off energy (the electron affinity) and is shown as decreasing along the y-axis. We now have
one mole of Cs cations and one mole of F anions. These ions combine to produce solid cesium fluoride. The enthalpy change in
this step is the negative of the lattice energy, so it is also an exothermic quantity. The total energy involved in this conversion is
equal to the experimentally determined enthalpy of formation, ΔH , of the compound from its elements. In this case, the overall
∘
f
change is exothermic.
Hess’s law can also be used to show the relationship between the enthalpies of the individual steps and the enthalpy of formation.
Table 7.3.1 shows this for cesium fluoride, CsF.
7.3.4 https://chem.libretexts.org/@go/page/119789
Table 7.3.1: Enthalpies of Select Transitions
1 1
One-half of the bond energy of F2 F (g) ⟶ F(g)
2
ΔH = D = 79 kJ/mol
2 2
1
∘ ∘
ΔH = ΔH = ΔHs + D + I E + (−EA) + (−ΔHlattic e )
f
Enthalpy of formation of CsF(s), add steps 1–5 1
2
Thus, the lattice energy can be calculated from other values. For cesium chloride, using this data, the lattice energy is:
ΔHlattice = (411 + 109 + 122 + 496 + 368) kJ = 770 kJ (7.3.4)
The Born-Haber cycle may also be used to calculate any one of the other quantities in the equation for lattice energy, provided that
the remainder is known. For example, if the relevant enthalpy of sublimation ΔH , ionization energy (IE), bond dissociation∘
s
enthalpy (D), lattice energy ΔHlattice, and standard enthalpy of formation ΔH are known, the Born-Haber cycle can be used to
∘
f
Summary
The strength of a covalent bond is measured by its bond dissociation energy, that is, the amount of energy required to break that
particular bond in a mole of molecules. Multiple bonds are stronger than single bonds between the same atoms. The enthalpy of a
reaction can be estimated based on the energy input required to break bonds and the energy released when new bonds are formed.
For ionic bonds, the lattice energy is the energy required to separate one mole of a compound into its gas phase ions. Lattice energy
increases for ions with higher charges and shorter distances between ions. Lattice energies are often calculated using the Born-
Haber cycle, a thermochemical cycle including all of the energetic steps involved in converting elements into an ionic compound.
Key Equations
Lattice energy for a solid MX: MX(s) ⟶ M n+
(g) + X
n−
(g) ΔHlattice
+ −
C(Z )(Z )
Lattice energy for an ionic crystal: ΔH lattice =
Ro
Footnotes
1. This question is taken from the Chemistry Advanced Placement Examination and is used with the permission of the Educational
Testing Service.
Glossary
Born-Haber cycle
thermochemical cycle relating the various energetic steps involved in the formation of an ionic solid from the relevant elements
lattice energy (ΔHlattice)
7.3.5 https://chem.libretexts.org/@go/page/119789
energy required to separate one mole of an ionic solid into its component gaseous ions
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Fuse School, Open Educational Resource free of charge, under a Creative Commons License: Attribution-NonCommercial CC
BY-NC (View License Deed: https://creativecommons.org/licenses/by-nc/4.0/)
Sci Show is a division of Complexly and videos are free to stream for educational purposes.
TED-Ed’s commitment to creating lessons worth sharing is an extension of TED’s mission of spreading great ideas. Within
TED-Ed’s growing library of TED-Ed animations, you will find carefully curated educational videos, many of which represent
collaborations between talented educators and animators nominated through the TED-Ed website.
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
7.3: Ionic Bond Formation and Strength is shared under a CC BY license and was authored, remixed, and/or curated by LibreTexts.
7.3.6 https://chem.libretexts.org/@go/page/119789
7.3: Ionic Bond Formations and Strength (Problems)
PROBLEM 7.3.1
Using the bond energies in Table 7.3.1, determine the approximate enthalpy change for each of the following reactions:
a. H (g) + Br (g) ⟶ 2 HBr(g)
2 2
Answer a
−114 kJ
Answer b
30 kJ
Answer c
−1055 kJ
PROBLEM 7.3.2
Using the bond energies in Table 7.3.1, determine the approximate enthalpy change for each of the following reactions:
a. H C = C H (g) + H (g) ⟶ H CCH (g)
2 2 2 3 3
Answer a
-128 kJ
Answer b
-5175 kJ
Click here to see a video of the solution
Problem 7.3.2
PROBLEM 7.3.3
How does the bond energy of HCl differ from the standard enthalpy of formation of HCl(g)?
Answer
The enthalpy of formation is -431.6 kJ, while the bond energy of H-Cl is -432 kJ. They are practically the same.
PROBLEM 7.3.4
7.3.1 https://chem.libretexts.org/@go/page/119790
Using the standard enthalpy of formation data in Appendix G, show how the standard enthalpy of formation of HCl(g) can be
used to determine the bond energy.
Answer
1 1
∘ ∘
HCl(g) ⟶ H (g) + Cl (g) ΔH = −ΔH
2 2 1 f[HCl(g)]
2 2
1
∘ ∘
H (g) ⟶ H(g) ΔH = ΔH
2 2 f[H(g)]
2
1
∘ ∘
Cl (g) ⟶ Cl(g) ΔH = ΔH
2 3 f[Cl(g)]
2
––––––––––––––––––––––––––––––––––––––––––––––––––
∘ ∘ ∘ ∘
HCl(g) ⟶ H(g) + Cl(g) ΔH = ΔH + ΔH + ΔH
298 1 2 3
∘ ∘ ∘ ∘
DHCl = ΔH = ΔH + ΔH + ΔH (7.3.1)
298 f[HCl(g)] f[H(g)] f[Cl(g)]
= 431.6 kJ (7.3.3)
PROBLEM 7.3.5
Using the standard enthalpy of formation data in Appendix G, determine which bond is stronger: the S–F bond in SF4(g) or in
SF6(g)?
Answer
The S–F bond in SF4 is stronger.
PROBLEM 7.3.6
Complete the following Lewis structure by adding bonds (not atoms), and then indicate the longest bond:
A Lewis structure is shown that is missing its bonds. It shows a horizontal row of six carbon atoms, equally spaced. Three hydrogen atoms are drawn around the first carbon, two around the second, one above the
fifth, and two by the sixth.
Answer
A Lewis structure is shown. A carbon atom that is single bonded to three hydrogen atoms is bonded to a second carbon atom. The second carbon atom is single bonded to two hydrogen atoms. The second
carbon atom is single bonded to a third carbon atom that is triple bonded to a fourth carbon atom and single bonded to a fifth carbon atom. The fifth carbon atom is single bonded to a hydrogen atom and
double bonded to a sixth carbon atom that is single bonded to two hydrogen atoms.
PROBLEM 7.3.7
Use principles of atomic structure to answer each of the following:1
a. The radius of the Ca atom is 197 pm; the radius of the Ca2+ ion is 99 pm. Account for the difference.
b. The lattice energy of CaO(s) is –3460 kJ/mol; the lattice energy of K2O is –2240 kJ/mol. Account for the difference.
c. Given these ionization values, explain the difference between Ca and K with regard to their first and second ionization
energies.
K 419 3050
Ca 590 1140
d. The first ionization energy of Mg is 738 kJ/mol and that of Al is 578 kJ/mol. Account for this difference.
7.3.2 https://chem.libretexts.org/@go/page/119790
Answer a
When two electrons are removed from the valence shell, the Ca radius loses the outermost energy level and reverts to the
lower n = 3 level, which is much smaller in radius.
Answer b
The +2 charge on calcium pulls the oxygen much closer compared with K, thereby increasing the lattice energy relative to a
less charged ion.
Answer c
Removal of the 4s electron in Ca requires more energy than removal of the 4s electron in K because of the stronger attraction
of the nucleus and the extra energy required to break the pairing of the electrons. The second ionization energy for K
requires that an electron be removed from a lower energy level, where the attraction is much stronger from the nucleus for
the electron. In addition, energy is required to unpair two electrons in a full orbital. For Ca, the second ionization potential
requires removing only a lone electron in the exposed outer energy level.
Answer d
In Al, the removed electron is relatively unprotected and unpaired in a p orbital. The higher energy for Mg mainly reflects
the unpairing of the 2s electron.
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Feedback
Think one of the answers above is wrong? Let us know here.
7.3: Ionic Bond Formations and Strength (Problems) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.
7.3.3 https://chem.libretexts.org/@go/page/119790
CHAPTER OVERVIEW
Unit 8: Solutions and Phase Changes
Unit Objectives
Describe the processes represented by typical heating and cooling curves, and compute heat flows and enthalpy changes
accompanying these processes
Explain the construction and use of a typical phase diagram
Use phase diagrams to identify stable phases at given temperatures and pressures, and to describe phase transitions resulting
from changes in these properties
Describe the supercritical fluid phase of matter
Describe the effects of temperature and pressure on solubility
State Henry’s law and use it in calculations involving the solubility of a gas in a liquid
Explain the degrees of solubility possible for liquid-liquid solutions
Define the concentration units of mass percentage, volume percentage, mass-volume percentage, parts-per-million (ppm),
and parts-per-billion (ppb)
Perform computations relating a solution’s concentration and its components’ volumes and/or masses using these units\
Express concentrations of solution components using mole fraction and molality
Describe the effect of solute concentration on various solution properties (vapor pressure, boiling point, freezing point, and
osmotic pressure)
Perform calculations using the mathematical equations that describe these various colligative effects
Unit 8: Solutions and Phase Changes is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
1
8.1: Heating Curves and Phase Changes
Skills to Develop
Describe the processes represented by typical heating and cooling curves, and compute heat flows and enthalpy changes
accompanying these processes
Explain the construction and use of a typical phase diagram
Use phase diagrams to identify stable phases at given temperatures and pressures, and to describe phase transitions resulting
from changes in these properties
Describe the supercritical fluid phase of matter
q = mcΔT (8.1.1)
where m is the mass of the substance and c is its specific heat. The relation applies to matter being heated or cooled, but not
undergoing a change in state. When a substance being heated or cooled reaches a temperature corresponding to one of its phase
transitions, further gain or loss of heat is a result of diminishing or enhancing intermolecular attractions, instead of increasing or
decreasing molecular kinetic energies. While a substance is undergoing a change in state, its temperature remains constant. Figure
8.1.1 shows a typical heating curve.
Figure 8.1.1 : A typical heating curve for a substance depicts changes in temperature that result as the substance absorbs
increasing amounts of heat. Plateaus in the curve (regions of constant temperature) are exhibited when the substance undergoes
phase transitions.
Consider the example of heating a pot of water to boiling. A stove burner will supply heat at a roughly constant rate; initially, this
heat serves to increase the water’s temperature. When the water reaches its boiling point, the temperature remains constant despite
the continued input of heat from the stove burner. This same temperature is maintained by the water as long as it is boiling. If the
burner setting is increased to provide heat at a greater rate, the water temperature does not rise, but instead the boiling becomes
more vigorous (rapid). This behavior is observed for other phase transitions as well: For example, temperature remains constant
while the change of state is in progress.
Example 8.1.1 : Total Heat Needed to Change Temperature and Phase for a Substance
8.1.1 https://chem.libretexts.org/@go/page/119748
How much heat is required to convert 135 g of ice at −15 °C into water vapor at 120 °C?
Solution
The transition described involves the following steps:
1. Heat ice from −15 °C to 0 °C
2. Melt ice
3. Heat water from 0 °C to 100 °C
4. Boil water
5. Heat steam from 100 °C to 120 °C
The heat needed to change the temperature of a given substance (with no change in phase) is: q = m × c × ΔT (see previous
chapter on thermochemistry). The heat needed to induce a given change in phase is given by q = n × ΔH.
Using these equations with the appropriate values for specific heat of ice, water, and steam, and enthalpies of fusion and
vaporization, we have:
qtotal = (m ⋅ c ⋅ ΔT )ice + n ⋅ ΔHfus + (m ⋅ c ⋅ ΔT )water + n ⋅ ΔHvap + (m ⋅ c ⋅ ΔT )steam
1 mol
= (135 g ⋅ 2.09 J/g ⋅ °C ⋅ 15°C) + (135 ⋅ ⋅ 6.01 kJ/mol)
18.02 g
1 mol
+(135 g ⋅ 4.18 J/g ⋅ °C ⋅ 100°C) + (135 g ⋅ ⋅ 40.67 kJ/mol)
18.02 g
Converting the quantities in J to kJ permits them to be summed, yielding the total heat required:
Exercise 8.1.1
What is the total amount of heat released when 94.0 g water at 80.0 °C cools to form ice at −30.0 °C?
Answer
40.5 kJ
In the previous unit, the variation of a liquid’s equilibrium vapor pressure with temperature was described. Considering the
definition of boiling point, plots of vapor pressure versus temperature represent how the boiling point of the liquid varies with
pressure. Also described was the use of heating and cooling curves to determine a substance’s melting (or freezing) point. Making
such measurements over a wide range of pressures yields data that may be presented graphically as a phase diagram. A phase
diagram combines plots of pressure versus temperature for the liquid-gas, solid-liquid, and solid-gas phase-transition equilibria of a
substance. These diagrams indicate the physical states that exist under specific conditions of pressure and temperature, and also
provide the pressure dependence of the phase-transition temperatures (melting points, sublimation points, boiling points). A typical
phase diagram for a pure substance is shown in Figure 8.1.2.
8.1.2 https://chem.libretexts.org/@go/page/119748
Figure 8.1.2 : The physical state of a substance and its phase-transition temperatures are represented graphically in a phase
diagram.
To illustrate the utility of these plots, consider the phase diagram for water shown in Figure 8.1.3.
Figure 8.1.3 : The pressure and temperature axes on this phase diagram of water are not drawn to constant scale in order to
illustrate several important properties.
We can use the phase diagram to identify the physical state of a sample of water under specified conditions of pressure and
temperature. For example, a pressure of 50 kPa and a temperature of −10 °C correspond to the region of the diagram labeled “ice.”
Under these conditions, water exists only as a solid (ice). A pressure of 50 kPa and a temperature of 50 °C correspond to the
“water” region—here, water exists only as a liquid. At 25 kPa and 200 °C, water exists only in the gaseous state. Note that on the
H2O phase diagram, the pressure and temperature axes are not drawn to a constant scale in order to permit the illustration of several
important features as described here.
The curve BC in Figure 8.1.3 is the plot of vapor pressure versus temperature as described in the previous module of this chapter.
This “liquid-vapor” curve separates the liquid and gaseous regions of the phase diagram and provides the boiling point for water at
any pressure. For example, at 1 atm, the boiling point is 100 °C. Notice that the liquid-vapor curve terminates at a temperature of
374 °C and a pressure of 218 atm, indicating that water cannot exist as a liquid above this temperature, regardless of the pressure.
The physical properties of water under these conditions are intermediate between those of its liquid and gaseous phases. This
unique state of matter is called a supercritical fluid, a topic that will be described in the next section of this module.
The solid-vapor curve, labeled AB in Figure 8.1.3, indicates the temperatures and pressures at which ice and water vapor are in
equilibrium. These temperature-pressure data pairs correspond to the sublimation, or deposition, points for water. If we could zoom
in on the solid-gas line in Figure 8.1.2, we would see that ice has a vapor pressure of about 0.20 kPa at −10 °C. Thus, if we place a
frozen sample in a vacuum with a pressure less than 0.20 kPa, ice will sublime. This is the basis for the “freeze-drying” process
often used to preserve foods, such as the ice cream shown in Figure 8.1.4.
8.1.3 https://chem.libretexts.org/@go/page/119748
>
Figure 8.1.4 : Freeze-dried foods, like this ice cream, are dehydrated by sublimation at pressures below the triple point for water.
(credit: ʺlwaoʺ/Flickr)
The solid-liquid curve labeled BD shows the temperatures and pressures at which ice and liquid water are in equilibrium,
representing the melting/freezing points for water. Note that this curve exhibits a slight negative slope (greatly exaggerated for
clarity), indicating that the melting point for water decreases slightly as pressure increases. Water is an unusual substance in this
regard, as most substances exhibit an increase in melting point with increasing pressure. This behavior is partly responsible for the
movement of glaciers, like the one shown in Figure 8.1.5. The bottom of a glacier experiences an immense pressure due to its
weight that can melt some of the ice, forming a layer of liquid water on which the glacier may more easily slide.
Figure 8.1.5 : The immense pressures beneath glaciers result in partial melting to produce a layer of water that provides
lubrication to assist glacial movement. This satellite photograph shows the advancing edge of the Perito Moreno glacier in
Argentina. (credit: NASA)
The point of intersection of all three curves is labeled B in Figure 8.1.3. At the pressure and temperature represented by this point,
all three phases of water coexist in equilibrium. This temperature-pressure data pair is called the triple point. At pressures lower
than the triple point, water cannot exist as a liquid, regardless of the temperature.
8.1.4 https://chem.libretexts.org/@go/page/119748
Example 8.1.2 : Determining the State of Water
Using the phase diagram for water given in Figure 10.4.2, determine the state of water at the following temperatures and
pressures:
a. −10 °C and 50 kPa
b. 25 °C and 90 kPa
c. 50 °C and 40 kPa
d. 80 °C and 5 kPa
e. −10 °C and 0.3 kPa
f. 50 °C and 0.3 kPa
Solution
Using the phase diagram for water, we can determine that the state of water at each temperature and pressure given are as
follows: (a) solid; (b) liquid; (c) liquid; (d) gas; (e) solid; (f) gas.
Exercise 8.1.2
What phase changes can water undergo as the temperature changes if the pressure is held at 0.3 kPa? If the pressure is held at 50
kPa?
Answer:
At 0.3 kPa: s⟶ g at −58 °C. At 50 kPa: s⟶ l at 0 °C, l ⟶ g at 78 °C
Consider the phase diagram for carbon dioxide shown in Figure 8.1.6 as another example. The solid-liquid curve exhibits a
positive slope, indicating that the melting point for CO2 increases with pressure as it does for most substances (water being a
notable exception as described previously). Notice that the triple point is well above 1 atm, indicating that carbon dioxide cannot
exist as a liquid under ambient pressure conditions. Instead, cooling gaseous carbon dioxide at 1 atm results in its deposition into
the solid state. Likewise, solid carbon dioxide does not melt at 1 atm pressure but instead sublimes to yield gaseous CO2. Finally,
notice that the critical point for carbon dioxide is observed at a relatively modest temperature and pressure in comparison to water.
Figure 8.1.6 : A phase diagram for carbon dioxide is shown. The pressure axis is plotted on a logarithmic scale to accommodate
the large range of values.
8.1.5 https://chem.libretexts.org/@go/page/119748
Example 8.1.3 : Determining the State of Carbon Dioxide
Using the phase diagram for carbon dioxide shown in Figure 10.4.5, determine the state of CO2 at the following temperatures
and pressures:
a. −30 °C and 2000 kPa
b. −60 °C and 1000 kPa
c. −60 °C and 100 kPa
d. 20 °C and 1500 kPa
e. 0 °C and 100 kPa
f. 20 °C and 100 kPa
Solution
Using the phase diagram for carbon dioxide provided, we can determine that the state of CO2 at each temperature and pressure
given are as follows: (a) liquid; (b) solid; (c) gas; (d) liquid; (e) gas; (f) gas.
Exercise 8.1.3
Determine the phase changes carbon dioxide undergoes when its temperature is varied, thus holding its pressure constant at
1500 kPa? At 500 kPa? At what approximate temperatures do these phase changes occur?
Answer
at 1500 kPa: s⟶ l at −45 °C, l⟶ g at −10 °C; at 500 kPa: s⟶ g at −58 °C
Supercritical Fluids
supercritical uids
liquefy a gas at its critical temperature is called the critical pressure. The critical temperatures and critical pressures of some
common substances are given in Table 8.1.1.
Table 8.1.1: Critical Temperatures and Critical Pressures of select substances
Substance Critical Temperature (K) Critical Pressure (atm)
8.1.6 https://chem.libretexts.org/@go/page/119748
Substance Critical Temperature (K) Critical Pressure (atm)
Supercritical CO2
Video 8.1.3 : The liquid to supercritical fluid transition for carbon dioxide.
Like a gas, a supercritical fluid will expand and fill a container, but its density is much greater than typical gas densities, typically
being close to those for liquids. Similar to liquids, these fluids are capable of dissolving nonvolatile solutes. They exhibit
essentially no surface tension and very low viscosities, however, so they can more effectively penetrate very small openings in a
solid mixture and remove soluble components. These properties make supercritical fluids extremely useful solvents for a wide
range of applications. For example, supercritical carbon dioxide has become a very popular solvent in the food industry, being used
to decaffeinate coffee, remove fats from potato chips, and extract flavor and fragrance compounds from citrus oils. It is nontoxic,
relatively inexpensive, and not considered to be a pollutant. After use, the CO2 can be easily recovered by reducing the pressure
and collecting the resulting gas.
Figure 8.1.7 : (a) A sealed container of liquid carbon dioxide slightly below its critical point is heated, resulting in (b) the
formation of the supercritical fluid phase. Cooling the supercritical fluid lowers its temperature and pressure below the critical
point, resulting in the reestablishment of separate liquid and gaseous phases (c and d). Colored floats illustrate differences in
density between the liquid, gaseous, and supercritical fluid states. (credit: modification of work by “mrmrobin”/YouTube)
8.1.7 https://chem.libretexts.org/@go/page/119748
If we shake a carbon dioxide fire extinguisher on a cool day (18 °C), we can hear liquid CO2 sloshing around inside the cylinder.
However, the same cylinder appears to contain no liquid on a hot summer day (35 °C). Explain these observations.
Solution
On the cool day, the temperature of the CO2 is below the critical temperature of CO2, 304 K or 31 °C (Table 8.1.1), so liquid
CO2 is present in the cylinder. On the hot day, the temperature of the CO2 is greater than its critical temperature of 31 °C. Above
this temperature no amount of pressure can liquefy CO2 so no liquid CO2 exists in the fire extinguisher.
Exercise 8.1.4
Ammonia can be liquefied by compression at room temperature; oxygen cannot be liquefied under these conditions. Why do the
two gases exhibit different behavior?
Answer
The critical temperature of ammonia is 405.5 K, which is higher than room temperature. The critical temperature of oxygen
is below room temperature; thus oxygen cannot be liquefied at room temperature.
Figure 8.1.8 : (a) Caffeine molecules have both polar and nonpolar regions, making it soluble in solvents of varying polarities. (b)
The schematic shows a typical decaffeination process involving supercritical carbon dioxide.
Supercritical fluid extraction using carbon dioxide is now being widely used as a more effective and environmentally friendly
decaffeination method (Figure 8.1.8). At temperatures above 304.2 K and pressures above 7376 kPa, CO2 is a supercritical fluid,
with properties of both gas and liquid. Like a gas, it penetrates deep into the coffee beans; like a liquid, it effectively dissolves
certain substances. Supercritical carbon dioxide extraction of steamed coffee beans removes 97−99% of the caffeine, leaving
coffee’s flavor and aroma compounds intact. Because CO2 is a gas under standard conditions, its removal from the extracted coffee
8.1.8 https://chem.libretexts.org/@go/page/119748
beans is easily accomplished, as is the recovery of the caffeine from the extract. The caffeine recovered from coffee beans via this
process is a valuable product that can be used subsequently as an additive to other foods or drugs.
Summary
Key Equations
−Δ Hv ap /RT
P = Ae
ΔHvap
ln P = − + ln A
RT
P2 ΔHvap 1 1
ln( ) = ( − )
P1 R T1 T2
Glossary
critical point
temperature and pressure above which a gas cannot be condensed into a liquid
phase diagram
pressure-temperature graph summarizing conditions under which the phases of a substance can exist
supercritical fluid
substance at a temperature and pressure higher than its critical point; exhibits properties intermediate between those of gaseous
and liquid states
triple point
temperature and pressure at which the vapor, liquid, and solid phases of a substance are in equilibrium
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
8.1.9 https://chem.libretexts.org/@go/page/119748
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Crash Course Physics: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
8.1: Heating Curves and Phase Changes is shared under a CC BY license and was authored, remixed, and/or curated by LibreTexts.
8.1.10 https://chem.libretexts.org/@go/page/119748
8.1: Heating Curves and Phase Changes (Problems)
PROBLEM 8.1.1
From the phase diagram for water, determine the state of water at:
a. 35 °C and 85 kPa
b. −15 °C and 40 kPa
c. −15 °C and 0.1 kPa
d. 75 °C and 3 kPa
e. 40 °C and 0.1 kPa
f. 60 °C and 50 kPa
Answer a
Liquid
Answer b
solid
Answer c
solid
Answer d
gas
Answer e
gas
Answer f
liquid
PROBLEM 8.1.2
Evaporation of sweat requires energy and thus take excess heat away from the body. Some of the water that you drink may
eventually be converted into sweat and evaporate. If you drink a 20-ounce bottle of water that had been in the refrigerator at
3.8 °C, how much heat is needed to convert all of that water into sweat and then to vapor?? (Note: Your body temperature is
36.6 °C. For the purpose of solving this problem, assume that the thermal properties of sweat are the same as for water.)
Answer
567g
20 ounces water= 567 g = 31.5 moles
18.0g/mole
Heating 567 g from 3.8 °C to 100 °C q1= m Cp ΔT = 567g x 4.18 J/g-°C x 96.2°C = 228 kJ
Vaporizing step q2= n ΔH v ap = 31.5 moles x 30.67 kJ/mole =1281 kJ
Cooling the vapor from 100 °C to 38.6 °C q3= m Cp ΔT =567g x 1.84 x (-61.4) = -64 kJ
qTotal = q1 + q2 + q3 = 228kJ + 1281kJ -64kJ = 1445 kJ
8.1.1 https://chem.libretexts.org/@go/page/119749
PROBLEM 8.1.3
How much heat is required to convert 422 g of liquid H2O at 23.5 °C into steam at 150 °C?
Answer
1130 kJ
Click here to see a video of the solution
Problem 8.1.3
PROBLEM 8.1.4
Titanium tetrachloride, TiCl4, has a melting point of −23.2 °C and has a ΔH fusion = 9.37 kJ/mol.
a. How much energy is required to melt 263.1 g TiCl4?
b. For TiCl4, which will likely have the larger magnitude: ΔH fusion or ΔH vaporization? Explain your reasoning.
Answer a
13.0 kJ
Answer b
It is likely that the heat of vaporization will have a larger magnitude since in the case of vaporization the intermolecular
interactions have to be completely overcome, while melting weakens or destroys only some of them.
PROBLEM 8.1.5
8.1.2 https://chem.libretexts.org/@go/page/119749
What phase changes will take place when water is subjected to varying pressure at a constant temperature of 0.005 °C? At 40
°C? At −40 °C?
Answer
At 0.005 °C vapor to solid to liquid. The triple point is at +0.01 °C and the freezing point is ever so slightly lower at 0.00
°C. So if one raises the pressure at a constant 0.005 °C one passes just to the left of the triple point and the vapor will first
solidify at 0.612 KPa and then the solid will melt and become a liquid between the 0.00 °C melting point and 0.612 kPa
At 40 °C vapor to liquid.
At - 40 °C vapor to solid
PROBLEM 8.1.6
8.1.3 https://chem.libretexts.org/@go/page/119749
From the phase diagram for carbon dioxide, determine the state of CO2 at:
a. 20 °C and 1000 kPa
b. 10 °C and 2000 kPa
c. 10 °C and 100 kPa
d. −40 °C and 500 kPa
e. −80 °C and 1500 kPa
f. −80 °C and 10 kPa
Answer a
liquid
Answer b
liquid
Answer c
liquid
Answer d
liquid
Answer e
solid
Answer f
8.1.4 https://chem.libretexts.org/@go/page/119749
solid
PROBLEM 8.1.7
Determine the phase changes that carbon dioxide undergoes as the pressure changes if the temperature is held at −50 °C? If the
temperature is held at −40 °C? At 20 °C?
Answer
−50 °C: gas to liquid to solid
−40 °C: gas to liquid to solid
20 °C: gas to liquid to solid
PROBLEM 8.1.8
8.1.5 https://chem.libretexts.org/@go/page/119749
Dry ice, CO2(s), does not melt at atmospheric pressure. It sublimes at a temperature of −78 °C. What is the lowest pressure at
which CO2(s) will melt to give CO2(l)? At approximately what temperature will this occur?
Answer
Dry ice, CO2(s), will melt to give CO2(l) at 5.11 atm at −56.6 °C, the triple point of carbon dioxide.
PROBLEM 8.1.9
Elemental carbon has one gas phase, one liquid phase, and three different solid phases, as shown in the phase diagram:
8.1.6 https://chem.libretexts.org/@go/page/119749
a. On the phase diagram, label the gas and liquid regions.
b. Graphite is the most stable phase of carbon at normal conditions. On the phase diagram, label the graphite phase.
c. If graphite at normal conditions is heated to 2500 K while the pressure is increased to 1010 Pa, it is converted into diamond.
Label the diamond phase.
d. Circle each triple point on the phase diagram.
e. In what phase does carbon exist at 5000 K and 108 Pa?
f. If the temperature of a sample of carbon increases from 3000 K to 5000 K at a constant pressure of 106 Pa, which phase
transition occurs, if any?
Answer a
Answer b
Answer c
8.1.7 https://chem.libretexts.org/@go/page/119749
Answer d
Answer e
liquid
Answer f
sublimation
Feedback
Think one of the answers above is wrong? Let us know here.
8.1: Heating Curves and Phase Changes (Problems) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.
8.1.8 https://chem.libretexts.org/@go/page/119749
8.2: Solubility and Intermolecular Forces
Skills to Develop
Describe the effects of temperature and pressure on solubility
State Henry’s law and use it in calculations involving the solubility of a gas in a liquid
Explain the degrees of solubility possible for liquid-liquid solutions
Imagine adding a small amount of salt to a glass of water, stirring until all the salt has dissolved, and then adding a bit more. You
can repeat this process until the salt concentration of the solution reaches its natural limit, a limit determined primarily by the
relative strengths of the solute-solute, solute-solvent, and solvent-solvent attractive forces discussed in the previous two modules of
this chapter. You can be certain that you have reached this limit because, no matter how long you stir the solution, undissolved salt
remains. The concentration of salt in the solution at this point is known as its solubility.
The solubility of a solute in a particular solvent is the maximum concentration that may be achieved under given conditions when
the dissolution process is at equilibrium. Referring to the example of salt in water:
+ −
NaCl(s) ⇌ Na (aq) + Cl (aq) (8.2.1)
When a solute’s concentration is equal to its solubility, the solution is said to be saturated with that solute. If the solute’s
concentration is less than its solubility, the solution is said to be unsaturated. A solution that contains a relatively low concentration
of solute is called dilute, and one with a relatively high concentration is called concentrated.
If we add more salt to a saturated solution of salt, we see it fall to the bottom and no more seems to dissolve. In fact, the added salt
does dissolve, as represented by the forward direction of the dissolution equation. Accompanying this process, dissolved salt will
precipitate, as depicted by the reverse direction of the equation. The system is said to be at equilibrium when these two reciprocal
processes are occurring at equal rates, and so the amount of undissolved and dissolved salt remains constant. Support for the
simultaneous occurrence of the dissolution and precipitation processes is provided by noting that the number and sizes of the
undissolved salt crystals will change over time, though their combined mass will remain the same.
Video 8.2.1 : Watch this impressive video showing the precipitation of sodium acetate from a supersaturated solution.
Solutions may be prepared in which a solute concentration exceeds its solubility. Such solutions are said to be supersaturated, and
they are interesting examples of nonequilibrium states. For example, the carbonated beverage in an open container that has not yet
“gone flat” is supersaturated with carbon dioxide gas; given time, the CO2 concentration will decrease until it reaches its
equilibrium value.
8.2.1 https://chem.libretexts.org/@go/page/119750
structure, note that the solubility of oxygen in the liquid hydrocarbon hexane, C6H14, is approximately 20 times greater than it is in
water.
Other factors also affect the solubility of a given substance in a given solvent. Temperature is one such factor, with gas solubility
typically decreasing as temperature increases (Figure 8.2.1). This is one of the major impacts resulting from the thermal pollution
of natural bodies of water.
Figure 8.2.1 : The solubilities of these gases in water decrease as the temperature increases. All solubilities were measured with a
constant pressure of 101.3 kPa (1 atm) of gas above the solutions.
When the temperature of a river, lake, or stream is raised abnormally high, usually due to the discharge of hot water from some
industrial process, the solubility of oxygen in the water is decreased. Decreased levels of dissolved oxygen may have serious
consequences for the health of the water’s ecosystems and, in severe cases, can result in large-scale fish kills (Figure 8.2.2).
Figure 8.2.2 : (a) The small bubbles of air in this glass of chilled water formed when the water warmed to room temperature and
the solubility of its dissolved air decreased. (b) The decreased solubility of oxygen in natural waters subjected to thermal pollution
can result in large-scale fish kills. (credit a: modification of work by Liz West; credit b: modification of work by U.S. Fish and
Wildlife Service)
The solubility of a gaseous solute is also affected by the partial pressure of solute in the gas to which the solution is exposed. Gas
solubility increases as the pressure of the gas increases. Carbonated beverages provide a nice illustration of this relationship. The
carbonation process involves exposing the beverage to a relatively high pressure of carbon dioxide gas and then sealing the
beverage container, thus saturating the beverage with CO2 at this pressure. When the beverage container is opened, a familiar hiss
is heard as the carbon dioxide gas pressure is released, and some of the dissolved carbon dioxide is typically seen leaving solution
in the form of small bubbles (Figure 8.2.3). At this point, the beverage is supersaturated with carbon dioxide and, with time, the
dissolved carbon dioxide concentration will decrease to its equilibrium value and the beverage will become “flat.”
8.2.2 https://chem.libretexts.org/@go/page/119750
Figure 8.2.3 : Opening the bottle of carbonated beverage reduces the pressure of the gaseous carbon dioxide above the beverage.
The solubility of CO2 is thus lowered, and some dissolved carbon dioxide may be seen leaving the solution as small gas bubbles.
(credit: modification of work by Derrick Coetzee)
For many gaseous solutes, the relation between solubility, Cg, and partial pressure, Pg, is a proportional one:
Cg = kPg (8.2.2)
where k is a proportionality constant that depends on the identities of the gaseous solute and solvent, and on the solution
temperature. This is a mathematical statement of Henry’s law: The quantity of an ideal gas that dissolves in a definite volume of
liquid is directly proportional to the pressure of the gas.
Cg
k =
Pg
−3 −1
1.38 × 10 mol L
=
101.3 kPa
−5 −1 −1
= 1.36 × 10 mol L kPa
−6 −1 −1
(1.82 × 10 mol L torr )
−5 −1 −1
1.36 × 10 mol L kPa × 20.7 kPa (8.2.4)
−6 −1 −1
(or 1.82 × 10 mol L torr × 155 torr)
−4 −1
= 2.82 × 10 mol L
Note that various units may be used to express the quantities involved in these sorts of computations. Any combination of units
that yield to the constraints of dimensional analysis are acceptable.
Exercise 8.2.1
8.2.3 https://chem.libretexts.org/@go/page/119750
Exposing a 100.0 mL sample of water at 0 °C to an atmosphere containing a gaseous solute at 20.26 kPa (152 torr) resulted in
the dissolution of 1.45 × 10−3 g of the solute. Use Henry’s law to determine the solubility of this gaseous solute when its
pressure is 101.3 kPa (760 torr).
Answer
7.25 × 10−3 g in 100.0 mL or 0.0725 g/L
Figure 8.2.4 : (a) US Navy divers undergo training in a recompression chamber. (b) Divers receive hyperbaric oxygen therapy.
As the diver ascends to the surface of the water, the ambient pressure decreases and the dissolved gases becomes less soluble. If
the ascent is too rapid, the gases escaping from the diver’s blood may form bubbles that can cause a variety of symptoms
ranging from rashes and joint pain to paralysis and death. To avoid DCS, divers must ascend from depths at relatively slow
speeds (10 or 20 m/min) or otherwise make several decompression stops, pausing for several minutes at given depths during the
ascent. When these preventive measures are unsuccessful, divers with DCS are often provided hyperbaric oxygen therapy in
pressurized vessels called decompression (or recompression) chambers (Figure 8.2.4).
Deviations from Henry’s law are observed when a chemical reaction takes place between the gaseous solute and the solvent. Thus,
for example, the solubility of ammonia in water does not increase as rapidly with increasing pressure as predicted by the law
because ammonia, being a base, reacts to some extent with water to form ammonium ions and hydroxide ions.
Gases can form supersaturated solutions. If a solution of a gas in a liquid is prepared either at low temperature or under pressure (or
both), then as the solution warms or as the gas pressure is reduced, the solution may become supersaturated. In 1986, more than
1700 people in Cameroon were killed when a cloud of gas, almost certainly carbon dioxide, bubbled from Lake Nyos (Figure
8.2.5), a deep lake in a volcanic crater. The water at the bottom of Lake Nyos is saturated with carbon dioxide by volcanic activity
beneath the lake. It is believed that the lake underwent a turnover due to gradual heating from below the lake, and the warmer, less-
dense water saturated with carbon dioxide reached the surface. Consequently, tremendous quantities of dissolved CO2 were
released, and the colorless gas, which is denser than air, flowed down the valley below the lake and suffocated humans and animals
living in the valley.
8.2.4 https://chem.libretexts.org/@go/page/119750
Figure 8.2.5 : (a) It is believed that the 1986 disaster that killed more than 1700 people near Lake Nyos in Cameroon resulted
when a large volume of carbon dioxide gas was released from the lake. (b) A CO2 vent has since been installed to help outgas the
lake in a slow, controlled fashion and prevent a similar catastrophe from happening in the future. (credit a: modification of work by
Jack Lockwood; credit b: modification of work by Bill Evans)
Figure 8.2.6 : Water and antifreeze are miscible; mixtures of the two are homogeneous in all proportions. (credit:
“dno1967”/Wikimedia commons)
Liquids that mix with water in all proportions are usually polar substances or substances that form hydrogen bonds. For such
liquids, the dipole-dipole attractions (or hydrogen bonding) of the solute molecules with the solvent molecules are at least as strong
as those between molecules in the pure solute or in the pure solvent. Hence, the two kinds of molecules mix easily. Likewise,
nonpolar liquids are miscible with each other because there is no appreciable difference in the strengths of solute-solute, solvent-
solvent, and solute-solvent intermolecular attractions. The solubility of polar molecules in polar solvents and of nonpolar molecules
in nonpolar solvents is, again, an illustration of the chemical axiom “like dissolves like.”
Two liquids that do not mix to an appreciable extent are called immiscible. Layers are formed when we pour immiscible liquids
into the same container. Gasoline, oil (Figure 8.2.7), benzene, carbon tetrachloride, some paints, and many other nonpolar liquids
are immiscible with water. The attraction between the molecules of such nonpolar liquids and polar water molecules is ineffectively
weak. The only strong attractions in such a mixture are between the water molecules, so they effectively squeeze out the molecules
of the nonpolar liquid. The distinction between immiscibility and miscibility is really one of degrees, so that miscible liquids are of
infinite mutual solubility, while liquids said to be immiscible are of very low (though not zero) mutual solubility.
8.2.5 https://chem.libretexts.org/@go/page/119750
Figure 8.2.7 : Water and oil are immiscible. Mixtures of these two substances will form two separate layers with the less dense oil
floating on top of the water. (credit: “Yortw”/Flickr)
Two liquids, such as bromine and water, that are of moderate mutual solubility are said to be partially miscible. Two partially
miscible liquids usually form two layers when mixed. In the case of the bromine and water mixture, the upper layer is water,
saturated with bromine, and the lower layer is bromine saturated with water. Since bromine is nonpolar, and, thus, not very soluble
in water, the water layer is only slightly discolored by the bright orange bromine dissolved in it. Since the solubility of water in
bromine is very low, there is no noticeable effect on the dark color of the bromine layer (Figure 8.2.8).
Figure 8.2.8 : Bromine (the deep orange liquid on the left) and water (the clear liquid in the middle) are partially miscible. The
top layer in the mixture on the right is a saturated solution of bromine in water; the bottom layer is a saturated solution of water in
bromine. (credit: Paul Flowers)
8.2.6 https://chem.libretexts.org/@go/page/119750
Figure 8.2.9 : This graph shows how the solubility of several solids changes with temperature.
The temperature dependence of solubility can be exploited to prepare supersaturated solutions of certain compounds. A solution
may be saturated with the compound at an elevated temperature (where the solute is more soluble) and subsequently cooled to a
lower temperature without precipitating the solute. The resultant solution contains solute at a concentration greater than its
equilibrium solubility at the lower temperature (i.e., it is supersaturated) and is relatively stable. Precipitation of the excess solute
can be initiated by adding a seed crystal (see the video in the Link to Learning earlier in this module) or by mechanically agitating
the solution. Some hand warmers, such as the one pictured in Figure 8.2.10, take advantage of this behavior.
Figure 8.2.10 : This hand warmer produces heat when the sodium acetate in a supersaturated solution precipitates. Precipitation
of the solute is initiated by a mechanical shockwave generated when the flexible metal disk within the solution is “clicked.” (credit:
modification of work by “Velela”/Wikimedia Commons)
Video 8.2.2 : This video shows the crystallization process occurring in a hand warmer.
Why Don't Oil and Water Mix?
8.2.7 https://chem.libretexts.org/@go/page/119750
Why don't oil and water mix? - John Pol…
Pol…
Video 8.2.3 : A look into why oil and water don't mix.
Summary
Key Equations
Cg = kPg
Glossary
Henry’s law
law stating the proportional relationship between the concentration of dissolved gas in a solution and the partial pressure of the
gas in contact with the solution
immiscible
of negligible mutual solubility; typically refers to liquid substances
8.2.8 https://chem.libretexts.org/@go/page/119750
miscible
mutually soluble in all proportions; typically refers to liquid substances
partially miscible
of moderate mutual solubility; typically refers to liquid substances
saturated
of concentration equal to solubility; containing the maximum concentration of solute possible for a given temperature and
pressure
solubility
extent to which a solute may be dissolved in water, or any solvent
supersaturated
of concentration that exceeds solubility; a nonequilibrium state
unsaturated
of concentration less than solubility
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
TED-Ed’s commitment to creating lessons worth sharing is an extension of TED’s mission of spreading great ideas. Within
TED-Ed’s growing library of TED-Ed animations, you will find carefully curated educational videos, many of which represent
collaborations between talented educators and animators nominated through the TED-Ed website.
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
8.2: Solubility and Intermolecular Forces is shared under a CC BY license and was authored, remixed, and/or curated by LibreTexts.
8.2.9 https://chem.libretexts.org/@go/page/119750
8.2: Solubility and Intermolecular Forces (Problems)
PROBLEM 8.2.1
Suggest an explanation for the observations that ethanol, C2H5OH, is completely miscible with water and that ethanethiol,
C2H5SH, is soluble only to the extent of 1.5 g per 100 mL of water.
Answer
The hydrogen bonds between water and C2H5OH are much stronger than the intermolecular attractions between water and
C2H5SH.
PROBLEM 8.2.2
Which of the following gases is expected to be most soluble in water? Explain your reasoning.
a. CH4
b. CCl4
c. CHCl3
Answer
(c) CHCl3 is expected to be most soluble in water. Of the three gases, only this one is polar and thus capable of experiencing
relatively strong dipole-dipole attraction to water molecules.
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Feedback
Think one of the answers above is wrong? Let us know here.
8.2: Solubility and Intermolecular Forces (Problems) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.
8.2.1 https://chem.libretexts.org/@go/page/119751
8.3: Concentrations of Solutions
Skills to Develop
Define the concentration units of mass percentage, volume percentage, mass-volume percentage, parts-per-million (ppm),
and parts-per-billion (ppb)
Perform computations relating a solution’s concentration and its components’ volumes and/or masses using these units\
Express concentrations of solution components using mole fraction and molality
Describe the effect of solute concentration on various solution properties (vapor pressure, boiling point, freezing point, and
osmotic pressure)
Perform calculations using the mathematical equations that describe these various colligative effects
Last term, we introduced molarity, a very useful measurement unit for evaluating the concentration of solutions. However, molarity
is only one measure of concentration. In this section, we will introduce some other units of concentration that are commonly used
in various applications, either for convenience or by convention.
Mass Percentage
Last term, we also introduced percent composition as a measure of the relative amount of a given element in a compound.
Percentages are also commonly used to express the composition of mixtures, including solutions. The mass percentage of a
solution component is defined as the ratio of the component’s mass to the solution’s mass, expressed as a percentage:
mass of component
mass percentage = × 100% (8.3.1)
mass of solution
We are generally most interested in the mass percentages of solutes, but it is also possible to compute the mass percentage of
solvent.
Mass percentage is also referred to by similar names such as percent mass, percent weight, weight/weight percent, and other
variations on this theme. The most common symbol for mass percentage is simply the percent sign, %, although more detailed
symbols are often used including %mass, %weight, and (w/w)%. Use of these more detailed symbols can prevent confusion of
mass percentages with other types of percentages, such as volume percentages (to be discussed later in this section).
Mass percentages are popular concentration units for consumer products. The label of a typical liquid bleach bottle (Figure 8.3.1)
cites the concentration of its active ingredient, sodium hypochlorite (NaOCl), as being 7.4%. A 100.0-g sample of bleach would
therefore contain 7.4 g of NaOCl.
Figure 8.3.1 : Liquid bleach is an aqueous solution of sodium hypochlorite (NaOCl). This brand has a concentration of 7.4%
NaOCl by mass.
8.3.1 https://chem.libretexts.org/@go/page/119752
1 g
3.75 mg glucose ×
1000 mg
% glucose = = 0.075% (8.3.2)
5.0 g spinal f luid
The computed mass percentage agrees with our rough estimate (it’s a bit less than 0.1%).
Note that while any mass unit may be used to compute a mass percentage (mg, g, kg, oz, and so on), the same unit must be used
for both the solute and the solution so that the mass units cancel, yielding a dimensionless ratio. In this case, we converted the
units of solute in the numerator from mg to g to match the units in the denominator. We could just as easily have converted the
denominator from g to mg instead. As long as identical mass units are used for both solute and solution, the computed mass
percentage will be correct.
Exercise 8.3.1
A bottle of a tile cleanser contains 135 g of HCl and 775 g of water. What is the percent by mass of HCl in this cleanser?
Answer
14.8%
For proper unit cancellation, the 0.500-L volume is converted into 500 mL, and the mass percentage is expressed as a ratio, 37.2
g HCl/g solution:
1.19 g solution 37.2 g HCl
500 mL solution ( )( ) = 221 g HCl (8.3.3)
mL solution 100 g solution
This mass of HCl is consistent with our rough estimate of approximately 200 g.
Exercise 8.3.2
What volume of concentrated HCl solution contains 125 g of HCl?
Answer
282 mL
Volume Percentage
Liquid volumes over a wide range of magnitudes are conveniently measured using common and relatively inexpensive laboratory
equipment. The concentration of a solution formed by dissolving a liquid solute in a liquid solvent is therefore often expressed as a
volume percentage, %vol or (v/v)%:
volume solute
volume percentage = × 100% (8.3.4)
volume solution
8.3.2 https://chem.libretexts.org/@go/page/119752
Example 8.3.3 : Calculations using Volume Percentage
Rubbing alcohol (isopropanol) is usually sold as a 70%vol aqueous solution. If the density of isopropyl alcohol is 0.785 g/mL,
how many grams of isopropyl alcohol are present in a 355 mL bottle of rubbing alcohol?
Solution
Per the definition of volume percentage, the isopropanol volume is 70% of the total solution volume. Multiplying the
isopropanol volume by its density yields the requested mass:
70 mL isopropyl alcohol 0.785 g isopropyl alcohol
355 mL solution( )( ) = 195 g isopropyl alcohol (8.3.5)
100 mL solution 1 mL isopropyl alcohol
Exercise 8.3.3
Wine is approximately 12% ethanol (CH CH OH ) by volume. Ethanol has a molar mass of 46.06 g/mol and a density 0.789
3 2
g/mL. How many moles of ethanol are present in a 750-mL bottle of wine?
Answer
1.5 mol ethanol
Mass-Volume Percentage
“Mixed” percentage units, derived from the mass of solute and the volume of solution, are popular for certain biochemical and
medical applications. A mass-volume percent is a ratio of a solute’s mass to the solution’s volume expressed as a percentage. The
specific units used for solute mass and solution volume may vary, depending on the solution. For example, physiological saline
solution, used to prepare intravenous fluids, has a concentration of 0.9% mass/volume (m/v), indicating that the composition is 0.9
g of solute per 100 mL of solution. The concentration of glucose in blood (commonly referred to as “blood sugar”) is also typically
expressed in terms of a mass-volume ratio. Though not expressed explicitly as a percentage, its concentration is usually given in
milligrams of glucose per deciliter (100 mL) of blood (Figure 8.3.2).
Figure 8.3.2 : “Mixed” mass-volume units are commonly encountered in medical settings. (a) The NaCl concentration of
physiological saline is 0.9% (m/v). (b) This device measures glucose levels in a sample of blood. The normal range for glucose
concentration in blood (fasting) is around 70–100 mg/dL. (credit a: modification of work by “The National Guard”/Flickr; credit
b: modification of work by Biswarup Ganguly).
mass solute 9
ppb = × 10 ppb (8.3.7)
mass solution
8.3.3 https://chem.libretexts.org/@go/page/119752
Both ppm and ppb are convenient units for reporting the concentrations of pollutants and other trace contaminants in water.
Concentrations of these contaminants are typically very low in treated and natural waters, and their levels cannot exceed relatively
low concentration thresholds without causing adverse effects on health and wildlife. For example, the EPA has identified the
maximum safe level of fluoride ion in tap water to be 4 ppm. Inline water filters are designed to reduce the concentration of
fluoride and several other trace-level contaminants in tap water (Figure 8.3.3).
Figure 8.3.3 : (a) In some areas, trace-level concentrations of contaminants can render unfiltered tap water unsafe for drinking
and cooking. (b) Inline water filters reduce the concentration of solutes in tap water. (credit a: modification of work by Jenn
Durfey; credit b: modification of work by “vastateparkstaff”/Wikimedia commons).
Example 8.3.4 : Parts per Million and Parts per Billion Concentrations
According to the EPA, when the concentration of lead in tap water reaches 15 ppb, certain remedial actions must be taken. What
is this concentration in ppm? At this concentration, what mass of lead (μg) would be contained in a typical glass of water (300
mL)?
Solution
The definitions of the ppm and ppb units may be used to convert the given concentration from ppb to ppm. Comparing these
two unit definitions shows that ppm is 1000 times greater than ppb (1 ppm = 103 ppb). Thus:
1 ppm
15 ppb × = 0.015 ppm (8.3.8)
3
10 ppb
The definition of the ppb unit may be used to calculate the requested mass if the mass of the solution is provided. However, only
the volume of solution (300 mL) is given, so we must use the density to derive the corresponding mass. We can assume the
density of tap water to be roughly the same as that of pure water (~1.00 g/mL), since the concentrations of any dissolved
substances should not be very large. Rearranging the equation defining the ppb unit and substituting the given quantities yields:
mass solute
9
ppb = × 10 ppb (8.3.9)
mass solution
1.00 g
15 ppb × 300 mL ×
mL −6
mass solute = = 4.5 × 10 g (8.3.11)
9
10 ppb
−6
1 μg
4.5 × 10 g× = 4.5 μg (8.3.12)
−6
10 g
Exercise 8.3.4
8.3.4 https://chem.libretexts.org/@go/page/119752
A 50.0-g sample of industrial wastewater was determined to contain 0.48 mg of mercury. Express the mercury concentration of
the wastewater in ppm and ppb units.
Answer
9.6 ppm, 9600 ppb
The properties of a solution are different from those of either the pure solute(s) or solvent. Many solution properties are dependent
upon the chemical identity of the solute. Compared to pure water, a solution of hydrogen chloride is more acidic, a solution of
ammonia is more basic, a solution of sodium chloride is more dense, and a solution of sucrose is more viscous. There are a few
solution properties, however, that depend only upon the total concentration of solute species, regardless of their identities. These
colligative properties include vapor pressure lowering, boiling point elevation, freezing point depression, and osmotic pressure.
This small set of properties is of central importance to many natural phenomena and technological applications, as will be
described in this module.
Because solution volumes vary with temperature, molar concentrations will likewise vary. When expressed as molarity, the
concentration of a solution with identical numbers of solute and solvent species will be different at different temperatures, due to
the contraction/expansion of the solution. More appropriate for calculations involving many colligative properties are mole-based
concentration units whose values are not dependent on temperature. Two such units are mole fraction (introduced in the previous
chapter on gases) and molality.
The mole fraction, X, of a component is the ratio of its molar amount to the total number of moles of all solution components:
mol A
XA = (8.3.14)
total mol of all components
Molality is a concentration unit defined as the ratio of the numbers of moles of solute to the mass of the solvent in kilograms:
mol solute
m = (8.3.15)
kg solvent
Since these units are computed using only masses and molar amounts, they do not vary with temperature and, thus, are better suited
for applications requiring temperature-independent concentrations, including several colligative properties, as will be described in
this chapter module.
water)?
Solution
(a) The mole fraction of ethylene glycol may be computed by first deriving molar amounts of both solution components and
then substituting these amounts into the unit definition.
1 mol C2 H4 (OH)2
mol C2 H4 (OH)2 = 2220 g × = 35.8 mol C2 H4 (OH)2
62.07 g C2 H4 (OH)2
1 mol H2 O
mol H2 O = 2000 g × = 111 mol H2 O
18.02 g H2 O
8.3.5 https://chem.libretexts.org/@go/page/119752
35.8 mol C2 H4 (OH)2
Xethylene glycol = = 0.245
(35.8 + 111) mol total
Notice that mole fraction is a dimensionless property, being the ratio of properties with identical units (moles).
(b) To find molality, we need to know the moles of the solute and the mass of the solvent (in kg).
First, use the given mass of ethylene glycol and its molar mass to find the moles of solute:
mol C2 H2 (OH)2
2220 g C2 H4 (OH)2 ( ) = 35.8 mol C2 H4 (OH)2 (8.3.16)
62.07 g
Exercise 8.3.5
What are the mole fraction and molality of a solution that contains 0.850 g of ammonia, NH3, dissolved in 125 g of water?
Answer
7.14 × 10−3; 0.399 m
The numerator for this solution’s mole fraction is, therefore, 3.0 mol NaCl. The denominator may be computed by deriving the
molar amount of water corresponding to 1.0 kg
1000 g mol H2 O
1.0 kg H2 O ( )( ) = 55 mol H2 O
1 kg 18.02 g
and then substituting these molar amounts into the definition for mole fraction.
8.3.6 https://chem.libretexts.org/@go/page/119752
mol H2 O
XH O =
2
mol NaCl + mol H2 O
55 mol H2 O
XH2 O =
3.0 mol NaCl + 55 mol H2 O
XH O = 0.95
2
mol NaCl
XNaCl =
mol NaCl + mol H2 O
XNaCl = 0.052
Exercise 8.3.6
The mole fraction of iodine, I , dissolved in dichloromethane, C H
2 2 C l2 , is 0.115. What is the molal concentration, m, of iodine
in this solution?
Answer
1.50 m
Dissolving a nonvolatile substance in a volatile liquid results in a lowering of the liquid’s vapor pressure. This phenomenon can be
rationalized by considering the effect of added solute molecules on the liquid's vaporization and condensation processes. To
vaporize, solvent molecules must be present at the surface of the solution. The presence of solute decreases the surface area
available to solvent molecules and thereby reduces the rate of solvent vaporization. Since the rate of condensation is unaffected by
the presence of solute, the net result is that the vaporization-condensation equilibrium is achieved with fewer solvent molecules in
the vapor phase (i.e., at a lower vapor pressure) (Figure 8.3.4). While this kinetic interpretation is useful, it does not account for
several important aspects of the colligative nature of vapor pressure lowering. A more rigorous explanation involves the property of
entropy, a topic of discussion in a later text chapter on thermodynamics. For purposes of understanding the lowering of a liquid's
vapor pressure, it is adequate to note that the greater entropy of a solution in comparison to its separate solvent and solute serves to
effectively stabilize the solvent molecules and hinder their vaporization. A lower vapor pressure results, and a correspondingly
higher boiling point as described in the next section of this module.
Figure 8.3.4 : The presence of nonvolatile solutes lowers the vapor pressure of a solution by impeding the evaporation of solvent
molecules.
The relationship between the vapor pressures of solution components and the concentrations of those components is described by
Raoult’s law: The partial pressure exerted by any component of an ideal solution is equal to the vapor pressure of the pure
component multiplied by its mole fraction in the solution.
8.3.7 https://chem.libretexts.org/@go/page/119752
∘
PA = XA P (8.3.23)
A
where PA is the partial pressure exerted by component A in the solution, P is the vapor pressure of pure A, and XA is the mole
A
∘
fraction of A in the solution. (Mole fraction is a concentration unit introduced in the chapter on gases.)
where
Kb is the boiling point elevation constant, or the ebullioscopic constant and
m is the molal concentration (molality) of all solute species.
Boiling point elevation constants are characteristic properties that depend on the identity of the solvent. Values of Kb for several
solvents are listed in Table 8.3.1.
Table 8.3.1: Boiling Point Elevation and Freezing Point Depression Constants for Several Solvents
Boiling Point (°C at 1 Freezing Point (°C at 1
Solvent Kb (Cm−1) Kf (Cm−1)
atm) atm)
The extent to which the vapor pressure of a solvent is lowered and the boiling point is elevated depends on the total number of
solute particles present in a given amount of solvent, not on the mass or size or chemical identities of the particles. A 1 m aqueous
solution of sucrose (342 g/mol) and a 1 m aqueous solution of ethylene glycol (62 g/mol) will exhibit the same boiling point
because each solution has one mole of solute particles (molecules) per kilogram of solvent.
Add the boiling point elevation to the pure solvent’s boiling point.
Boiling temperature = 80.1 °C + 0.83 °C = 80.9 °C
Exercise 8.3.7
What is the boiling point of the antifreeze described in Example 8.3.4?
8.3.8 https://chem.libretexts.org/@go/page/119752
Answer
109.2 °C
1. Convert from grams to moles of I2 using the molar mass of I2 in the unit conversion factor.
Result: 0.363 mol
2. Determine the molality of the solution from the number of moles of solute and the mass of solvent, in kilograms.
Result: 0.454 m
3. Use the direct proportionality between the change in boiling point and molal concentration to determine how much the
boiling point changes.
Result: 1.65 °C
4. Determine the new boiling point from the boiling point of the pure solvent and the change.
Result: 62.91 °C
Exercise 8.3.8
What is the boiling point of a solution of 1.0 g of glycerin, C3H5(OH)3, in 47.8 g of water? Assume an ideal solution.
Answer
100.12 °C
Distillation of Solutions
Distillation is a technique for separating the components of mixtures that is widely applied in both in the laboratory and in
industrial settings. It is used to refine petroleum, to isolate fermentation products, and to purify water. This separation technique
involves the controlled heating of a sample mixture to selectively vaporize, condense, and collect one or more components of
interest. A typical apparatus for laboratory-scale distillations is shown in Figure 8.3.5.
8.3.9 https://chem.libretexts.org/@go/page/119752
Figure 8.3.5 : A typical laboratory distillation unit is shown in (a) a photograph and (b) a schematic diagram of the components.
(credit a: modification of work by “Rifleman82”/Wikimedia commons; credit b: modification of work by “Slashme”/Wikimedia
Commons)
Oil refineries use large-scale fractional distillation to separate the components of crude oil. The crude oil is heated to high
temperatures at the base of a tall fractionating column, vaporizing many of the components that rise within the column. As
vaporized components reach adequately cool zones during their ascent, they condense and are collected. The collected liquids are
simpler mixtures of hydrocarbons and other petroleum compounds that are of appropriate composition for various applications
(e.g., diesel fuel, kerosene, gasoline), as depicted in Figure 8.3.6.
Figure 8.3.6 : Crude oil is a complex mixture that is separated by large-scale fractional distillation to isolate various simpler
mixtures.
automobile radiators. Seawater freezes at a lower temperature than fresh water, and so the Arctic and Antarctic oceans remain
unfrozen even at temperatures below 0 °C (as do the body fluids of fish and other cold-blooded sea animals that live in these
oceans).
8.3.10 https://chem.libretexts.org/@go/page/119752
Figure 8.3.7 : Rock salt (NaCl), calcium chloride (CaCl2), or a mixture of the two are used to melt ice. (credit: modification of
work by Eddie Welker)
The decrease in freezing point of a dilute solution compared to that of the pure solvent, ΔTf, is called the freezing point depression
and is directly proportional to the molal concentration of the solute
ΔTf = Kf m (8.3.25)
where
m is the molal concentration of the solute in the solvent and
Kf is called the freezing point depression constant (or cryoscopic constant).
Just as for boiling point elevation constants, these are characteristic properties whose values depend on the chemical identity of the
solvent. Values of Kf for several solvents are listed in Table 8.3.1.
2. Subtract the freezing point change observed from the pure solvent’s freezing point. [\mathrm{Freezing\: Temperature=5.5\:
°C−1.7\:°C=3.8\:°C}\]
Exercise 8.3.9
What is the freezing point of a 1.85 m solution of a nonvolatile nonelectrolyte solute in nitrobenzene?
Answer
−9.3 °C
8.3.11 https://chem.libretexts.org/@go/page/119752
elevate the boiling point, making the fluid useful in both winter and summer. Heated glycols are often sprayed onto the surface
of airplanes prior to takeoff in inclement weather in the winter to remove ice that has already formed and prevent the formation
of more ice, which would be particularly dangerous if formed on the control surfaces of the aircraft (Video 8.3.1).
Video 8.3.1 : Freezing point depression is exploited to remove ice from the control surfaces of aircraft.
Figure 8.3.8 : These phase diagrams show water (solid curves) and an aqueous solution of nonelectrolyte (dashed curves).
The liquid-vapor curve for the solution is located beneath the corresponding curve for the solvent, depicting the vapor pressure
lowering, ΔP, that results from the dissolution of nonvolatile solute. Consequently, at any given pressure, the solution’s boiling
point is observed at a higher temperature than that for the pure solvent, reflecting the boiling point elevation, ΔTb, associated with
the presence of nonvolatile solute. The solid-liquid curve for the solution is displaced left of that for the pure solvent, representing
the freezing point depression, ΔTf, that accompanies solution formation. Finally, notice that the solid-gas curves for the solvent and
its solution are identical. This is the case for many solutions comprising liquid solvents and nonvolatile solutes. Just as for
vaporization, when a solution of this sort is frozen, it is actually just the solvent molecules that undergo the liquid-to-solid
transition, forming pure solid solvent that excludes solute species. The solid and gaseous phases, therefore, are composed solvent
only, and so transitions between these phases are not subject to colligative effects.
8.3.12 https://chem.libretexts.org/@go/page/119752
Example 8.3.10 : Determining Molar Mass from Freezing Point Depression
A solution of 4.00 g of a nonelectrolyte dissolved in 55.0 g of benzene is found to freeze at 2.32 °C. What is the molar mass of
this compound?
Solution
We can solve this problem using the following steps.
1. Determine the change in freezing point from the observed freezing point and the freezing point of pure benzene (Table
11.5.1).
ΔTf = 5.5 °C − 2.32 °C = 3.2 °C
2. Determine the molal concentration from Kf, the freezing point depression constant for benzene (Table 11.5.1), and ΔTf.
ΔTf = Kf m
ΔTf 3.2 °C
m = = = 0.63 m
−1
Kf 5.12 °Cm
3. Determine the number of moles of compound in the solution from the molal concentration and the mass of solvent used to
make the solution.
0.62 mol solute
Moles of solute = × 0.0550 kg solvent = 0.035 mol
1.00 kg solvent
4. Determine the molar mass from the mass of the solute and the number of moles in that mass.
4.00 g 2
Molar mass = = 1.2 × 10 g/mol
0.034 mol
Exercise 8.3.10
A solution of 35.7 g of a nonelectrolyte in 220.0 g of chloroform has a boiling point of 64.5 °C. What is the molar mass of this
compound?
Answer
1.8 × 102 g/mol
1. Convert the osmotic pressure to atmospheres, then determine the molar concentration from the osmotic pressure.
\[\Pi=\mathrm{\dfrac{5.9\:torr×1\:atm}{760\:torr}=7.8×10^{−3}\:atm}\]
Π = M RT (8.3.27)
8.3.13 https://chem.libretexts.org/@go/page/119752
−3
Π 7.8 × 10 atm
−4
M = = = 3.2 × 10 M
RT (0.08206 L atm/mol K)(295 K)
2. Determine the number of moles of hemoglobin in the solution from the concentration and the volume of the solution.
−4
3.2 × 10 mol
−4
moles of hemoglobin = × 0.500 L solution = 1.6 × 10 mol
1 L solution
3. Determine the molar mass from the mass of hemoglobin and the number of moles in that mass.
10.0 g
4
molar mass = = 6.2 × 10 g/mol
−4
1.6 × 10 mol
Exercise 8.3.11
What is the molar mass of a protein if a solution of 0.02 g of the protein in 25.0 mL of solution has an osmotic pressure of 0.56
torr at 25 °C?
Answer
2.7 × 104 g/mol
Convert from grams to moles of NaCl using the molar mass of NaCl in the unit conversion factor. Result: 0.072 mol NaCl
Determine the number of moles of ions present in the solution using the number of moles of ions in 1 mole of NaCl as the
conversion factor (2 mol ions/1 mol NaCl). Result: 0.14 mol ions
Determine the molality of the ions in the solution from the number of moles of ions and the mass of solvent, in kilograms.
Result: 1.1 m
Use the direct proportionality between the change in freezing point and molal concentration to determine how much the
freezing point changes. Result: 2.0 °C
Determine the new freezing point from the freezing point of the pure solvent and the change. Result: −2.0 °C
Check each result as a self-assessment.
Exercise 8.3.13
8.3.14 https://chem.libretexts.org/@go/page/119752
Assume that each of the ions in calcium chloride, CaCl2, has the same effect on the freezing point of water as a nonelectrolyte
molecule. Calculate the freezing point of a solution of 0.724 g of CaCl2 in 175 g of water.
Answer
−0.208 °C
Assuming complete dissociation, a 1.0 m aqueous solution of NaCl contains 2.0 mole of ions (1.0 mol Na+ and 1.0 mol Cl−) per
each kilogram of water, and its freezing point depression is expected to be
ΔTf = 2.0 mol ions/kg water × 1.86 °C kg water/mol ion = 3.7 °C. (8.3.28)
When this solution is actually prepared and its freezing point depression measured, however, a value of 3.4 °C is obtained. Similar
discrepancies are observed for other ionic compounds, and the differences between the measured and expected colligative property
values typically become more significant as solute concentrations increase. These observations suggest that the ions of sodium
chloride (and other strong electrolytes) are not completely dissociated in solution.
To account for this and avoid the errors accompanying the assumption of total dissociation, an experimentally measured parameter
named in honor of Nobel Prize-winning German chemist Jacobus Henricus van’t Hoff is used. The van’t Hoff factor (i) is defined
as the ratio of solute particles in solution to the number of formula units dissolved:
moles of particles in solution
i = (8.3.29)
moles of formula units dissolved
Values for measured van’t Hoff factors for several solutes, along with predicted values assuming complete dissociation, are shown
in Table 8.3.2.
Table 8.3.2: Expected and Observed van’t Hoff Factors for Several 0.050 m Aqueous Electrolyte Solutions
Electrolyte Particles in Solution i (Predicted) i (Measured)
MgSO4 Mg2+, SO 2−
4
2 1.3
In 1923, the chemists Peter Debye and Erich Hückel proposed a theory to explain the apparent incomplete ionization of strong
electrolytes. They suggested that although interionic attraction in an aqueous solution is very greatly reduced by solvation of the
ions and the insulating action of the polar solvent, it is not completely nullified. The residual attractions prevent the ions from
behaving as totally independent particles (Figure 8.3.9). In some cases, a positive and negative ion may actually touch, giving a
solvated unit called an ion pair. Thus, the activity, or the effective concentration, of any particular kind of ion is less than that
indicated by the actual concentration. Ions become more and more widely separated the more dilute the solution, and the residual
interionic attractions become less and less. Thus, in extremely dilute solutions, the effective concentrations of the ions (their
activities) are essentially equal to the actual concentrations. Note that the van’t Hoff factors for the electrolytes in Table 8.3.2 are
for 0.05 m solutions, at which concentration the value of i for NaCl is 1.9, as opposed to an ideal value of 2.
8.3.15 https://chem.libretexts.org/@go/page/119752
>
Figure 8.3.9 : Ions become more and more widely separated the more dilute the solution, and the residual interionic attractions
become less.
Summary
Key Equations
ΔTb = Kbm
ΔTf = Kfm
Π = MRT
Footnotes
1. A nonelectrolyte shown for comparison.
8.3.16 https://chem.libretexts.org/@go/page/119752
Glossary
mass percentage
ratio of solute-to-solution mass expressed as a percentage
mass-volume percent
ratio of solute mass to solution volume, expressed as a percentage
volume percentage
ratio of solute-to-solution volume expressed as a percentage
colligative property
property of a solution that depends only on the concentration of a solute species
ion pair
solvated anion/cation pair held together by moderate electrostatic attraction
molality (m)
a concentration unit defined as the ratio of the numbers of moles of solute to the mass of the solvent in kilograms
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
8.3.17 https://chem.libretexts.org/@go/page/119752
8.3: Concentrations of Solutions is shared under a CC BY license and was authored, remixed, and/or curated by LibreTexts.
8.3.18 https://chem.libretexts.org/@go/page/119752
8.3: Concentrations of Solutions (Problems)
PROBLEM 8.3.1
What mass of a concentrated solution of nitric acid (68.0% HNO3 by mass) is needed to prepare 400.0 g of a 10.0% solution of
HNO3 by mass?
Answer
58.8 g
PROBLEM 8.3.2
What mass of a 4.00% NaOH solution by mass contains 15.0 g of NaOH?
Answer
375 g
Problem 8.3.2
PROBLEM 8.3.3
What mass of solid NaOH (97.0% NaOH by mass) is required to prepare 1.00 L of a 10.0% solution of NaOH by mass? The
density of the 10.0% solution is 1.109 g/mL.
Answer
114 g
PROBLEM 8.3.4
The hardness of water (hardness count) is usually expressed in parts per million (by mass) of CaCO , which is equivalent to
3
milligrams of CaCO per liter of water. What is the molar concentration of Ca2+ ions in a water sample with a hardness count
3
of 175 mg CaCO3/L?
Answer
−3
1.75 × 10 M
PROBLEM 8.3.5
A throat spray is 1.40% by mass phenol, C 6
H OH
5
, in water. If the solution has a density of 0.9956 g/mL, calculate the molarity
of the solution.
Answer
8.3.1 https://chem.libretexts.org/@go/page/119753
0.148 M
Problem 8.3.5
PROBLEM 8.3.6
Copper(I) iodide (CuI) is often added to table salt as a dietary source of iodine. How many moles of CuI are contained in 1.00 lb
(454 g) of table salt containing 0.0100% CuI by mass?
Answer
−4
2.38 × 10 mol
PROBLEM 8.3.7
What are the mole fractions of H3PO4 and water in a solution of 14.5 g of H3PO4 in 125 g of water?
Answer
XH3 PO4 = 0.021
XH O = 0.979
2
Problem 8.3.7
PROBLEM 8.3.8
What are the mole fractions of HNO3 and water in a concentrated solution of nitric acid (68.0% HNO3 by mass)?
Answer
XHNO3 = 0.378
8.3.2 https://chem.libretexts.org/@go/page/119753
XH O = 0.622
2
PROBLEM 8.3.9
Calculate the mole fraction of each solute and solvent:
a. 583 g of H2SO4 in 1.50 kg of water—the acid solution used in an automobile battery
b. 0.86 g of NaCl in 1.00 × 102 g of water—a solution of sodium chloride for intravenous injection
c. 46.85 g of codeine, C18H21NO3, in 125.5 g of ethanol, C2H5OH
d. 25 g of I2 in 125 g of ethanol, C2H5OH
Answer a
XH2 SO4 = 0.067
XH O = 0.933
2
Answer b
XHCl = 0.0026
XH2 O = 0.9974
Answer c
Xcodiene = 0.054
XEtOH = 0.946
Answer d
XI2 = 0.035
XEtOH = 0.965
Problem 8.3.9
PROBLEM 8.3.10
Calculate the mole fraction of each solute and solvent:
a. 0.710 kg of sodium carbonate (washing soda), Na2CO3, in 10.0 kg of water—a saturated solution at 0 °C
b. 125 g of NH4NO3 in 275 g of water—a mixture used to make an instant ice pack
c. 25 g of Cl2 in 125 g of dichloromethane, CH2Cl2
d. 0.372 g of histamine, C5H9N, in 125 g of chloroform, CHCl3
Answer a
XNa2 C O3 = 0.0119
8.3.3 https://chem.libretexts.org/@go/page/119753
XH O = 0.988
2
Answer b
XNH NO3 = 0.09927
4
XH2 O = 0.907
Answer c
XCl2 = 0.192
XC H C I2 = 0.808
2
Answer d
XC H9 N = 0.00426
5
XCHCl3 = 0.997
PROBLEM 8.3.11
What is the difference between a 1 M solution and a 1 m solution?
Answer
In a 1 M solution, the mole is contained in exactly 1 L of solution. In a 1 m solution, the mole is contained in exactly 1 kg of
solvent.
PROBLEM 8.3.12
What is the molality of phosphoric acid, H3PO4, in a solution of 14.5 g of H3PO4 in 125 g of water?
Answer
1.18 m
PROBLEM 8.3.13
What is the molality of nitric acid in a concentrated solution of nitric acid (68.0% HNO3 by mass)?
Answer
33.7 m
PROBLEM 8.3.14
Calculate the molality of each of the following solutions:
a. 0.710 kg of sodium carbonate (washing soda), Na2CO3, in 10.0 kg of water—a saturated solution at 0°C
b. 125 g of NH4NO3 in 275 g of water—a mixture used to make an instant ice pack
c. 25 g of Cl2 in 125 g of dichloromethane, CH2Cl2
d. 0.372 g of histamine, C5H9N, in 125 g of chloroform, CHCl3
Answer
6.70 × 10−1 m
Answer
5.67 m
Answer
2.8 m
8.3.4 https://chem.libretexts.org/@go/page/119753
Answer
0.0358 m
PROBLEM 8.3.15
A 13.0% solution of K2CO3 by mass has a density of 1.09 g/cm3. Calculate the molality of the solution.
Answer
1.08 m
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Feedback
Think one of the answers above is wrong? Let us know here.
8.3: Concentrations of Solutions (Problems) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by
LibreTexts.
8.3.5 https://chem.libretexts.org/@go/page/119753
CHAPTER OVERVIEW
Unit 9: Semiconductors
Skills to Develop
Outline the basic quantum-mechanical approach to deriving molecular orbitals from atomic orbitals
Describe traits of bonding and antibonding molecular orbitals
Calculate bond orders based on molecular electron configurations
Define and describe the bonding and properties of ionic, molecular, metallic, and covalent network crystalline solids
Describe the main types of crystalline solids: ionic solids, metallic solids, covalent network solids, and molecular solids
Explain the ways in which crystal defects can occur in a solid
Describe the general preparation, properties, and uses of the metalloids
Describe the preparation, properties, and compounds of boron and silicon
Explain the properties of some covalent-network solids
Describe semiconductors and some of their properties
Thumbnail: Silicon crystals are the most common semiconducting materials used in microelectronics and photovoltaics. Image
used with permission (CC BY 3.0; Jurii)
Unit 9: Semiconductors is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
1
9.1: Bond Types and Molecular Orbital Theory
Skills to Develop
Outline the basic quantum-mechanical approach to deriving molecular orbitals from atomic orbitals
Describe traits of bonding and antibonding molecular orbitals
Calculate bond orders based on molecular electron configurations
For almost every covalent molecule that exists, we can now draw the Lewis structure, predict the electron-pair geometry, predict
the molecular geometry, and come close to predicting bond angles. However, one of the most important molecules we know, the
oxygen molecule O2, presents a problem with respect to its Lewis structure. We would write the following Lewis structure for O2:
This electronic structure adheres to all the rules governing Lewis theory. There is an O=O double bond, and each oxygen atom has
eight electrons around it. However, this picture is at odds with the magnetic behavior of oxygen. By itself, O2 is not magnetic, but it
is attracted to magnetic fields. Thus, when we pour liquid oxygen past a strong magnet, it collects between the poles of the magnet
and defies gravity. Such attraction to a magnetic field is called paramagnetism, and it arises in molecules that have unpaired
electrons. And yet, the Lewis structure of O2 indicates that all electrons are paired. How do we account for this discrepancy?
Magnetic susceptibility measures the force experienced by a substance in a magnetic field. When we compare the weight of a
sample to the weight measured in a magnetic field (Figure 9.1.1), paramagnetic samples that are attracted to the magnet will appear
heavier because of the force exerted by the magnetic field. We can calculate the number of unpaired electrons based on the increase
in weight.
Figure 9.1.1 : A Gouy balance compares the mass of a sample in the presence of a magnetic field with the mass with the
electromagnet turned off to determine the number of unpaired electrons in a sample.
Experiments show that each O2 molecule has two unpaired electrons. The Lewis-structure model does not predict the presence of
these two unpaired electrons. Unlike oxygen, the apparent weight of most molecules decreases slightly in the presence of an
inhomogeneous magnetic field. Materials in which all of the electrons are paired are diamagnetic and weakly repel a magnetic
field. Paramagnetic and diamagnetic materials do not act as permanent magnets. Only in the presence of an applied magnetic field
do they demonstrate attraction or repulsion.
9.1.1 https://chem.libretexts.org/@go/page/119766
levitating frog
Video 9.1.1 : Water, like most molecules, contains all paired electrons. Living things contain a large percentage of water, so they
demonstrate diamagnetic behavior. If you place a frog near a sufficiently large magnet, it will levitate. You can see videos of
diamagnetic floating frogs, strawberries, and more (https://www.youtube.com/watch?v=A1vyB-O5i6E)
Molecular orbital theory (MO theory) provides an explanation of chemical bonding that accounts for the paramagnetism of the
oxygen molecule. It also explains the bonding in a number of other molecules, such as violations of the octet rule and more
molecules with more complicated bonding (beyond the scope of this text) that are difficult to describe with Lewis structures.
Additionally, it provides a model for describing the energies of electrons in a molecule and the probable location of these electrons.
Unlike valence bond theory, which uses hybrid orbitals that are assigned to one specific atom, MO theory uses the combination of
atomic orbitals to yield molecular orbitals that are delocalized over the entire molecule rather than being localized on its constituent
atoms. MO theory also helps us understand why some substances are electrical conductors, others are semiconductors, and still
others are insulators. Table 9.1.1 summarizes the main points of the two complementary bonding theories. Both theories provide
different, useful ways of describing molecular structure.
Table 9.1.1: Comparison of Bonding Theories
Valence Bond Theory Molecular Orbital Theory
considers bonds as localized between one pair of atoms considers electrons delocalized throughout the entire molecule
creates bonds from overlap of atomic orbitals (s, p, d…) and hybrid
combines atomic orbitals to form molecular orbitals (σ, σ*, π, π*)
orbitals (sp, sp2, sp3…)
Molecular orbital theory describes the distribution of electrons in molecules in much the same way that the distribution of electrons
in atoms is described using atomic orbitals. Using quantum mechanics, the behavior of an electron in a molecule is still described
by a wave function, Ψ, analogous to the behavior in an atom. Just like electrons around isolated atoms, electrons around atoms in
molecules are limited to discrete (quantized) energies. The region of space in which a valence electron in a molecule is likely to be
found is called a molecular orbital (Ψ2). Like an atomic orbital, a molecular orbital is full when it contains two electrons with
opposite spin.
We will consider the molecular orbitals in molecules composed of two identical atoms (H2 or Cl2, for example). Such molecules
are called homonuclear diatomic molecules. In these diatomic molecules, several types of molecular orbitals occur.
The mathematical process of combining atomic orbitals to generate molecular orbitals is called the linear combination of atomic
orbitals (LCAO). The wave function describes the wavelike properties of an electron. Molecular orbitals are combinations of
atomic orbital wave functions. Combining waves can lead to constructive interference, in which peaks line up with peaks, or
destructive interference, in which peaks line up with troughs (Figure 9.1.2). In orbitals, the waves are three dimensional, and they
9.1.2 https://chem.libretexts.org/@go/page/119766
combine with in-phase waves producing regions with a higher probability of electron density and out-of-phase waves producing
nodes, or regions of no electron density.
Figure 9.1.2 : (a) When in-phase waves combine, constructive interference produces a wave with greater amplitude. (b) When out-
of-phase waves combine, destructive interference produces a wave with less (or no) amplitude.
There are two types of molecular orbitals that can form from the overlap of two atomic s orbitals on adjacent atoms. The two types
are illustrated in Figure 8.4.3. The in-phase combination produces a lower energy σs molecular orbital (read as "sigma-s") in which
most of the electron density is directly between the nuclei. The out-of-phase addition (which can also be thought of as subtracting
the wave functions) produces a higher energy σ molecular orbital (read as "sigma-s-star") molecular orbital in which there is a
∗
s
node between the nuclei. The asterisk signifies that the orbital is an antibonding orbital. Electrons in a σs orbital are attracted by
both nuclei at the same time and are more stable (of lower energy) than they would be in the isolated atoms. Adding electrons to
these orbitals creates a force that holds the two nuclei together, so we call these orbitals bonding orbitals. Electrons in the σ ∗
s
orbitals are located well away from the region between the two nuclei. The attractive force between the nuclei and these electrons
pulls the two nuclei apart. Hence, these orbitals are called antibonding orbitals. Electrons fill the lower-energy bonding orbital
before the higher-energy antibonding orbital, just as they fill lower-energy atomic orbitals before they fill higher-energy atomic
orbitals.
Figure 9.1.3 : Sigma (σ) and sigma-star (σ*) molecular orbitals are formed by the combination of two s atomic orbitals. The plus
(+) signs indicate the locations of nuclei.
In p orbitals, the wave function gives rise to two lobes with opposite phases, analogous to how a two-dimensional wave has both
parts above and below the average. We indicate the phases by shading the orbital lobes different colors. When orbital lobes of the
same phase overlap, constructive wave interference increases the electron density. When regions of opposite phase overlap, the
destructive wave interference decreases electron density and creates nodes. When p orbitals overlap end to end, they create σ and
σ* orbitals (Figure 9.1.4). If two atoms are located along the x-axis in a Cartesian coordinate system, the two px orbitals overlap
end to end and form σpx (bonding) and σ (antibonding) (read as "sigma-p-x" and "sigma-p-x star," respectively). Just as with s-
∗
px
orbital overlap, the asterisk indicates the orbital with a node between the nuclei, which is a higher-energy, antibonding orbital.
Figure 9.1.4 : Combining wave functions of two p atomic orbitals along the internuclear axis creates two molecular orbitals, σp
and σ . ∗
p
9.1.3 https://chem.libretexts.org/@go/page/119766
The side-by-side overlap of two p orbitals gives rise to a pi (π) bonding molecular orbital and a π* antibonding molecular orbital, as
shown in Figure 9.1.5. In valence bond theory, we describe π bonds as containing a nodal plane containing the internuclear axis
and perpendicular to the lobes of the p orbitals, with electron density on either side of the node. In molecular orbital theory, we
describe the π orbital by this same shape, and a π bond exists when this orbital contains electrons. Electrons in this orbital interact
with both nuclei and help hold the two atoms together, making it a bonding orbital. For the out-of-phase combination, there are two
nodal planes created, one along the internuclear axis and a perpendicular one between the nuclei.
Figure 9.1.5 : Side-by-side overlap of each two p orbitals results in the formation of two π molecular orbitals. Combining the out-
of-phase orbitals results in an antibonding molecular orbital with two nodes. One contains the internuclear axis, and one is
perpendicular to the axis. Combining the in-phase orbitals results in a bonding orbital. There is a node (blue plane) containing the
internuclear axis with the two lobes of the orbital located above and below this node.
In the molecular orbitals of diatomic molecules, each atom also has two sets of p orbitals oriented side by side (py and pz), so these
four atomic orbitals combine pairwise to create two π orbitals and two π* orbitals. The πpy and π orbitals are oriented at right
∗
py
angles to the πpz and π orbitals. Except for their orientation, the πpy and πpz orbitals are identical and have the same energy; they
∗
pz
are degenerate orbitals. The π and π antibonding orbitals are also degenerate and identical except for their orientation. A total
∗
py
∗
pz
of six molecular orbitals results from the combination of the six atomic p orbitals in two atoms: σpx and σ , πpy and π , πpz and
∗
px
∗
py
π .
∗
pz
Example 9.1.1
Molecular Orbitals Predict what type (if any) of molecular orbital would result from adding the wave functions so each pair of
orbitals shown overlap. The orbitals are all similar in energy.
Solution
a. This is an in-phase combination, resulting in a σ3p orbital
b. This will not result in a new orbital because the in-phase component (bottom) and out-of-phase component (top) cancel out.
Only orbitals with the correct alignment can combine.
c. This is an out-of-phase combination, resulting in a π orbital.
∗
3p
Exercise 9.1.1
Label the molecular orbital shown as σ or π, bonding or antibonding and indicate where the node occurs.
9.1.4 https://chem.libretexts.org/@go/page/119766
Answer
The orbital is located along the internuclear axis, so it is a σ orbital. There is a node bisecting the internuclear axis, so it is an
antibonding orbital.
Figure 9.1.6 : The molecule shown, HIV-1 protease, is an important target for pharmaceutical research. By designing
molecules that bind to this protein, scientists are able to drastically inhibit the progress of the disease.
9.1.5 https://chem.libretexts.org/@go/page/119766
Figure 9.1.7 : This is the molecular orbital diagram for the homonuclear diatomic Be , showing the molecular orbitals of the
+
valence shell only. The molecular orbitals are filled in the same manner as atomic orbitals, using the Aufbau principle and Hund’s
rule.
We predict the distribution of electrons in these molecular orbitals by filling the orbitals in the same way that we fill atomic
orbitals, by the Aufbau principle. Lower-energy orbitals fill first, electrons spread out among degenerate orbitals before pairing,
and each orbital can hold a maximum of two electrons with opposite spins (Figure 9.1.7). Just as we write electron configurations
for atoms, we can write the molecular electronic configuration by listing the orbitals with superscripts indicating the number of
electrons present. For clarity, we place parentheses around molecular orbitals with the same energy. In this case, each orbital is at a
different energy, so parentheses separate each orbital. Thus we would expect a diatomic molecule or ion containing seven electrons
(such as Be ) would have the molecular electron configuration (σ ) (σ ) (σ ) (σ ) . It is common to omit the core electrons
+
2 1s
2 ∗
1s
2
2s
2 ∗
2s
1
from molecular orbital diagrams and configurations and include only the valence electrons.
Bond Order
The filled molecular orbital diagram shows the number of electrons in both bonding and antibonding molecular orbitals. The net
contribution of the electrons to the bond strength of a molecule is identified by determining the bond order that results from the
filling of the molecular orbitals by electrons.
When using Lewis structures to describe the distribution of electrons in molecules, we define bond order as the number of bonding
pairs of electrons between two atoms. Thus a single bond has a bond order of 1, a double bond has a bond order of 2, and a triple
bond has a bond order of 3. We define bond order differently when we use the molecular orbital description of the distribution of
electrons, but the resulting bond order is usually the same. The MO technique is more accurate and can handle cases when the
Lewis structure method fails, but both methods describe the same phenomenon.
In the molecular orbital model, an electron contributes to a bonding interaction if it occupies a bonding orbital and it contributes to
an antibonding interaction if it occupies an antibonding orbital. The bond order is calculated by subtracting the destabilizing
(antibonding) electrons from the stabilizing (bonding) electrons. Since a bond consists of two electrons, we divide by two to get the
bond order. We can determine bond order with the following equation:
(number of bonding electrons) − (number of antibonding electrons)
bond order = (9.1.1)
2
The order of a covalent bond is a guide to its strength; a bond between two given atoms becomes stronger as the bond order
increases. If the distribution of electrons in the molecular orbitals between two atoms is such that the resulting bond would have a
bond order of zero, a stable bond does not form. We next look at some specific examples of MO diagrams and bond orders.
9.1.6 https://chem.libretexts.org/@go/page/119766
Bonding in Diatomic Molecules
A dihydrogen molecule (H2) forms from two hydrogen atoms. When the atomic orbitals of the two atoms combine, the electrons
occupy the molecular orbital of lowest energy, the σ1s bonding orbital. A dihydrogen molecule, H2, readily forms because the
energy of a H2 molecule is lower than that of two H atoms. The σ1s orbital that contains both electrons is lower in energy than
either of the two 1s atomic orbitals.
A molecular orbital can hold two electrons, so both electrons in the H2 molecule are in the σ1s bonding orbital; the electron
configuration is (σ ) . We represent this configuration by a molecular orbital energy diagram (Figure 9.1.8) in which a single
1s
2
upward arrow indicates one electron in an orbital, and two (upward and downward) arrows indicate two electrons of opposite spin.
Figure 9.1.8 :The molecular orbital energy diagram predicts that H2 will be a stable molecule with lower energy than the
separated atoms.
PREVIEW: Band Theory in Extended Systems
When two identical atomic orbitals on different atoms combine, two molecular orbitals result (e.g., H in Figure 9.1.8). The
2
bonding orbital is lower in energy than the original atomic orbitals because the atomic orbitals are in-phase in the molecular
orbital. The antibonding orbital is higher in energy than the original atomic orbitals because the atomic orbitals are out-of-phase.
In a solid, similar things happen, but on a much larger scale. Remember that even in a small sample there are a huge number of
atoms (typically > 1023 atoms), and therefore a huge number of atomic orbitals that may be combined into molecular orbitals.
When N valence atomic orbitals, all of the same energy and each containing one (1) electron, are combined, N/2 (filled) bonding
orbitals and N/2 (empty) antibonding orbitals will result. Each bonding orbital will show an energy lowering as the atomic
orbitals are mostly in-phase, but each of the bonding orbitals will be a little different and have slightly different energies. The
antibonding orbitals will show an increase in energy as the atomic orbitals are mostly out-of-phase, but each of the antibonding
orbitals will also be a little different and have slightly different energies. The allowed energy levels for all the bonding orbitals
are so close together that they form a band, called the valence band. Likewise, all the antibonding orbitals are very close
together and form a band, called the conduction band. Figure 9.1.9) shows the bands for three important classes of materials:
insulators, semiconductors, and conductors.
Figure 9.1.9 : Molecular orbitals in solids are so closely spaced that they are described as bands. The valence band is lower in
energy and the conduction band is higher in energy. The type of solid is determined by the size of the “band gap” between the
valence and conduction bands. Only a very small amount of energy is required to move electrons from the valence band to the
conduction band in a conductor, and so they conduct electricity well. In an insulator, the band gap is large, so that very few
9.1.7 https://chem.libretexts.org/@go/page/119766
electrons move, and they are poor conductors of electricity. Semiconductors are in between: they conduct electricity better than
insulators, but not as well as conductors.
In order to conduct electricity, electrons must move from the filled valence band to the empty conduction band where they can
move throughout the solid. The size of the band gap, or the energy difference between the top of the valence band and the
bottom of the conduction band, determines how easy it is to move electrons between the bands. Only a small amount of energy
is required in a conductor because the band gap is very small. This small energy difference is “easy” to overcome, so they are
good conductors of electricity. In an insulator, the band gap is so “large” that very few electrons move into the conduction band;
as a result, insulators are poor conductors of electricity. Semiconductors conduct electricity when “moderate” amounts of energy
are provided to move electrons out of the valence band and into the conduction band. Semiconductors, such as silicon, are found
in many electronics.
Semiconductors are used in devices such as computers, smartphones, and solar cells. Solar cells produce electricity when light
provides the energy to move electrons out of the valence band. The electricity that is generated may then be used to power a
light or tool, or it can be stored for later use by charging a battery. As of December 2014, up to 46% of the energy in sunlight
could be converted into electricity using solar cells.
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
9.1: Bond Types and Molecular Orbital Theory is shared under a CC BY license and was authored, remixed, and/or curated by LibreTexts.
9.1.8 https://chem.libretexts.org/@go/page/119766
9.1: Bond Types and Molecular Orbital Theory (Problems)
PROBLEM 9.1.1
How are the following similar, and how do they differ?
(a) σ molecular orbitals and π molecular orbitals
(b) bonding orbitals and antibonding orbitals
Answer a
Similarities: Both are bonding orbitals that can contain a maximum of two electrons. Differences: σ orbitals are end-to-end
combinations of atomic orbitals, whereas π orbitals are formed by side-by-side overlap of orbitals
Answer b
Similarities: Both are orbitals that can contain two electrons. Differences: Bonding orbitals result in holding two or more
atoms together. Antibonding orbitals have the effect of destabilizing any bonding that has occurred.
PROBLEM 9.1.2
Can a molecule with an odd number of electrons ever be diamagnetic? Explain why or why not.
Answer
An odd number of electrons can never be paired, regardless of the arrangement of the molecular orbitals. It will always be
paramagnetic.
PROBLEM 9.1.3
Can a molecule with an even number of electrons ever be paramagnetic? Explain why or why not.
Answer
Yes, you could potentially have unpaired electrons in each orbital
PROBLEM 9.1.4
Why are bonding molecular orbitals lower in energy than the parent atomic orbitals?
Answer
Bonding orbitals have electron density in close proximity to more than one nucleus. The interaction between the bonding
positively charged nuclei and negatively charged electrons stabilizes the system.
PROBLEM 9.1.5
Explain why an electron in the bonding molecular orbital in the H2 molecule has a lower energy than an electron in the 1s
atomic orbital of either of the separated hydrogen atoms.
Answer
The pairing of the two bonding electrons lowers the energy of the system relative to the energy of the nonbonded electrons.
9.1.1 https://chem.libretexts.org/@go/page/119767
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Feedback
Think one of the answers above is wrong? Let us know here.
9.1: Bond Types and Molecular Orbital Theory (Problems) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or
curated by LibreTexts.
9.1.2 https://chem.libretexts.org/@go/page/119767
9.2: The Solid State of Matter
Skills to Develop
Define and describe the bonding and properties of ionic, molecular, metallic, and covalent network crystalline solids
Describe the main types of crystalline solids: ionic solids, metallic solids, covalent network solids, and molecular solids
Explain the ways in which crystal defects can occur in a solid
When most liquids are cooled, they eventually freeze and form crystalline solids, solids in which the atoms, ions, or molecules are
arranged in a definite repeating pattern. It is also possible for a liquid to freeze before its molecules become arranged in an orderly
pattern. The resulting materials are called amorphous solids or noncrystalline solids (or, sometimes, glasses). The particles of such
solids lack an ordered internal structure and are randomly arranged (Figure 9.2.1).
Figure 9.2.1 : The entities of a solid phase may be arranged in a regular, repeating pattern (crystalline solids) or randomly
(amorphous).
Metals and ionic compounds typically form ordered, crystalline solids. Substances that consist of large molecules, or a mixture of
molecules whose movements are more restricted, often form amorphous solids. For examples, candle waxes are amorphous solids
composed of large hydrocarbon molecules. Some substances, such as silicon dioxide (Figure 9.2.2), can form either crystalline or
amorphous solids, depending on the conditions under which it is produced. Also, amorphous solids may undergo a transition to the
crystalline state under appropriate conditions.
Figure 9.2.2 : (a) Silicon dioxide, SiO2, is abundant in nature as one of several crystalline forms of the mineral quartz. (b) Rapid
cooling of molten SiO2 yields an amorphous solid known as “fused silica”.
Crystalline solids are generally classified according the nature of the forces that hold its particles together. These forces are
primarily responsible for the physical properties exhibited by the bulk solids. The following sections provide descriptions of the
major types of crystalline solids: ionic, metallic, covalent network, and molecular.
9.2.1 https://chem.libretexts.org/@go/page/119768
Ionic Solids
Ionic solids, such as sodium chloride and nickel oxide, are composed of positive and negative ions that are held together by
electrostatic attractions, which can be quite strong (Figure 9.2.3). Many ionic crystals also have high melting points. This is due to
the very strong attractions between the ions—in ionic compounds, the attractions between full charges are (much) larger than those
between the partial charges in polar molecular compounds. This will be looked at in more detail in a later discussion of lattice
energies. Although they are hard, they also tend to be brittle, and they shatter rather than bend. Ionic solids do not conduct
electricity; however, they do conduct when molten or dissolved because their ions are free to move. Many simple compounds
formed by the reaction of a metallic element with a nonmetallic element are ionic.
Metallic Bonding
In the early 1900's, Paul Drüde came up with the "sea of electrons" metallic bonding theory by modeling metals as a mixture of
atomic cores (atomic cores = positive nuclei + inner shell of electrons) and valence electrons. Metallic bonds occur among metal
atoms. Whereas ionic bonds join metals to non-metals, metallic bonding joins a bulk of metal atoms. A sheet of aluminum foil and
a copper wire are both places where you can see metallic bonding in action.
Metals tend to have high melting points and boiling points suggesting strong bonds between the atoms. Even a soft metal like
sodium (melting point 97.8°C) melts at a considerably higher temperature than the element (neon) which precedes it in the Periodic
Table. Sodium has the electronic structure 1s22s22p63s1. When sodium atoms come together, the electron in the 3s atomic orbital of
one sodium atom shares space with the corresponding electron on a neighboring atom to form a molecular orbital - in much the
same sort of way that a covalent bond is formed.
The difference, however, is that each sodium atom is being touched by eight other sodium atoms - and the sharing occurs between
the central atom and the 3s orbitals on all of the eight other atoms. Each of these eight is in turn being touched by eight sodium
atoms, which in turn are touched by eight atoms - and so on and so on, until you have taken in all the atoms in that lump of sodium.
All of the 3s orbitals on all of the atoms overlap to give a vast number of molecular orbitals that extend over the whole piece of
metal. There have to be huge numbers of molecular orbitals, of course, because any orbital can only hold two electrons.
The electrons can move freely within these molecular orbitals, and so each electron becomes detached from its parent atom. The
electrons are said to be delocalized. The metal is held together by the strong forces of attraction between the positive nuclei and the
delocalized electrons (Figure 9.2.4).
Figure 9.2.4 : Metallic Bonding: The Electron Sea Model: Positive atomic nuclei (orange circles) surrounded by a sea of
delocalized electrons (yellow circles).
9.2.2 https://chem.libretexts.org/@go/page/119768
This is sometimes described as "an array of positive ions in a sea of electrons". If you are going to use this view, beware! Is a metal
made up of atoms or ions? It is made of atoms. Each positive center in the diagram represents all the rest of the atom apart from the
outer electron, but that electron has not been lost - it may no longer have an attachment to a particular atom, but it's still there in the
structure. Sodium metal is therefore written as Na, not Na . +
9.2.3 https://chem.libretexts.org/@go/page/119768
Figure 9.2.5 : The "sea of electrons" is free to flow about the crystal of positive metal ions. These flowing electron can conduct
electrical change when an electric field is applied (e.g., a battery). (CC-BY-SA; OpenStax and Rafaelgarcia).
Malleability and Ductility: The electron-sea model of metals not only explains their electrical properties but their
malleability and ductility as well. The sea of electrons surrounding the protons acts like a cushion, and so when the metal
is hammered on, for instance, the overall composition of the structure of the metal is not harmed or changed. The protons
may be rearranged but the sea of electrons with adjust to the new formation of protons and keep the metal intact. When one
layer of ions in an electron sea moves along one space with respect to the layer below it, the crystal structure does not
fracture but is only deformed (Figure 9.2.6).
Figure 9.2.6 : Malleability of metals originate from each of moving layer of atoms with respect to each other. The final
situation is much the same as the initial. Thus if we hit a metal with a hammer, the crystals do not shatter, but merely change
their shape, This is very different from the behavior of ionic crystals.
Heat capacity: This is explained by the ability of free electrons to move about the solid.
Luster: The free electrons can absorb photons in the "sea," so metals are opaque-looking. Electrons on the surface can
bounce back light at the same frequency that the light hits the surface, therefore the metal appears to be shiny.
However, these observations are only qualitative, and not quantitative, so they cannot be tested. The "Sea of Electrons" theory
stands today only as an oversimplified model of how metallic bonding works.
In a molten metal, the metallic bond is still present, although the ordered structure has been broken down. The metallic bond is not
fully broken until the metal boils. That means that boiling point is actually a better guide to the strength of the metallic bond than
melting point is. On melting, the bond is loosened, not broken. The strength of a metallic bond depends on three things:
1. The number of electrons that become delocalized from the metal
2. The charge of the cation (metal).
3. The size of the cation.
A strong metallic bond will be the result of more delocalized electrons, which causes the effective nuclear charge on electrons on
the cation to increase, in effect making the size of the cation smaller. Metallic bonds are strong and require a great deal of energy to
break, and therefore metals have high melting and boiling points. A metallic bonding theory must explain how so much bonding
can occur with such few electrons (since metals are located on the left side of the periodic table and do not have many electrons in
their valence shells). The theory must also account for all of a metal's unique chemical and physical properties.
9.2.4 https://chem.libretexts.org/@go/page/119768
“ionic” and “covalent” bonding are idealized concepts and most bonds exist on a two-dimensional continuum described by the van
Arkel-Ketelaar Triangle (Figure 9.2.7).
Figure 9.2.7 : van Arkel-Ketelaar Triangle plots the difference in electronegativity (Δχ) and the average electronegativity in a bond
(∑ χ ). the top region is where bonds are mostly ionic, the lower left region is where bonding is metallic, and the lower right region
is where the bonding is covalent.
Bond triangles or van Arkel–Ketelaar triangles (named after Anton Eduard van Arkel and J. A. A. Ketelaar) are triangles used for
showing different compounds in varying degrees of ionic, metallic and covalent bonding. In 1941 van Arkel recognized three
extreme materials and associated bonding types. Using 36 main group elements, such as metals, metalloids and non-metals, he
placed ionic, metallic and covalent bonds on the corners of an equilateral triangle, as well as suggested intermediate species. The
bond triangle shows that chemical bonds are not just particular bonds of a specific type. Rather, bond types are interconnected and
different compounds have varying degrees of different bonding character (for example, polar covalent bonds).
we can rate the dominant bond between the compounds. On the right side of Figure 9.2.7 (from ionic to covalent) should be
compounds with varying difference in electronegativity. The compounds with equal electronegativity, such as Cl (chlorine) are 2
placed in the covalent corner, while the ionic corner has compounds with large electronegativity difference, such as NaCl (table
salt). The bottom side (from metallic to covalent) contains compounds with varying degree of directionality in the bond. At one
extreme is metallic bonds with delocalized bonding and at the other are covalent bonds in which the orbitals overlap in a particular
direction. The left side (from ionic to metallic) is meant for delocalized bonds with varying electronegativity difference.
9.2.5 https://chem.libretexts.org/@go/page/119768
The Three Extremes in bonding
In general:
Metallic bonds have low Δχ and low average ∑ χ.
Ionic bonds have moderate-to-high Δχ and moderate values of average ∑ χ.
Covalent bonds have moderate to high average ∑ χ and can exist with moderately low Δχ.
Example 9.2.2
Use the tables of electronegativities (Table A2) and Figure 9.2.7 to estimate the following values
difference in electronegativity (Δχ)
average electronegativity in a bond (∑ χ)
percent ionic character
likely bond type
for the selected compounds:
a. AsH (e.g., in arsine AsH )
b. SrLi
c. KF .
Solution
a: AsH
The electronegativity of As is 2.18
The electronegativity of H is 2.22
Using Equations 9.2.1 and 9.2.2:
χA + χB
∑χ =
2
2.18 + 2.22
=
2
= 2.2
Δχ = χA − χB
= 2.18 − 2.22
= 0.04
From Figure 9.2.4, the bond is fairly nonpolar and has a low ionic character (10% or less)
The bonding is in the middle of a covalent bond and a metallic bond
b: SrLi
The electronegativity of Sr is 0.95
The electronegativity of Li is 0.98
Using Equations 9.2.1 and 9.2.2:
χA + χB
∑χ =
2
0.95 + 0.98
=
2
= 0.965
Δχ = χA − χB
= 0.98 − 0.95
= 0.025
9.2.6 https://chem.libretexts.org/@go/page/119768
From Figure 9.2.4, the bond is fairly nonpolar and has a low ionic character (~3% or less)
The bonding is likely metallic.
c: KF
The electronegativity of K is 0.82
The electronegativity of F is 3.98
Using Equations 9.2.1 and 9.2.2:
χA + χB
∑χ =
2
0.82 + 3.98
=
2
= 2.4
Δχ = χA − χB
= |0.82 − 3.98|
= 3.16
From Figure 9.2.4, the bond is fairly polar and has a high ionic character (~75%)
The bonding is likely ionic.
Exercise 9.2.2
Contrast the bonding of NaCl and silicon tetrafluoride.
Answer
NaCl is an ionic crystal structure, and an electrolyte when dissolved in water; Δχ = 1.58, average ∑ χ = 1.79 , while
silicon tetrafluoride is covalent (molecular, non-polar gas; Δχ = 2.08, average ∑ χ = 2.94.
Metallic Solids
Metallic solids such as crystals of copper, aluminum, and iron are formed by metal atoms Figure 9.2.8. The structure of metallic
crystals is often described as a uniform distribution of atomic nuclei within a “sea” of delocalized electrons. The atoms within such
a metallic solid are held together by a unique force known as metallic bonding that gives rise to many useful and varied bulk
properties. All exhibit high thermal and electrical conductivity, metallic luster, and malleability. Many are very hard and quite
strong. Because of their malleability (the ability to deform under pressure or hammering), they do not shatter and, therefore, make
useful construction materials. The melting points of the metals vary widely. Mercury is a liquid at room temperature, and the alkali
metals melt below 200 °C. Several post-transition metals also have low melting points, whereas the transition metals melt at
temperatures above 1000 °C. These differences reflect differences in strengths of metallic bonding among the metals.
9.2.7 https://chem.libretexts.org/@go/page/119768
atoms in these solids are held together by a network of covalent bonds, as shown in Figure 9.2.9. To break or to melt a covalent
network solid, covalent bonds must be broken. Because covalent bonds are relatively strong, covalent network solids are typically
characterized by hardness, strength, and high melting points. For example, diamond is one of the hardest substances known and
melts above 3500 °C.
Figure 9.2.9 . A covalent crystal contains a three-dimensional network of covalent bonds, as illustrated by the structures of
diamond, silicon dioxide, silicon carbide, and graphite. Graphite is an exceptional example, composed of planar sheets of covalent
crystals that are held together in layers by noncovalent forces. Unlike typical covalent solids, graphite is very soft and electrically
conductive.
Molecular Solids
Molecular solids, such as ice, sucrose (table sugar), and iodine, as shown in Figure 9.2.10, are composed of neutral molecules. The
strengths of the attractive forces between the units present in different crystals vary widely, as indicated by the melting points of the
crystals. Small symmetrical molecules (nonpolar molecules), such as H2, N2, O2, and F2, have weak attractive forces and form
molecular solids with very low melting points (below −200 °C). Substances consisting of larger, nonpolar molecules have larger
attractive forces and melt at higher temperatures. Molecular solids composed of molecules with permanent dipole moments (polar
molecules) melt at still higher temperatures. Examples include ice (melting point, 0 °C) and table sugar (melting point, 185 °C).
Figure 9.2.10 : Carbon dioxide (CO2) consists of small, nonpolar molecules and forms a molecular solid with a melting point of
−78 °C. Iodine (I2) consists of larger, nonpolar molecules and forms a molecular solid that melts at 114 °C.
Properties of Solids
A crystalline solid, like those listed in Table 9.2.1 has a precise melting temperature because each atom or molecule of the same
type is held in place with the same forces or energy. Thus, the attractions between the units that make up the crystal all have the
same strength and all require the same amount of energy to be broken. The gradual softening of an amorphous material differs
dramatically from the distinct melting of a crystalline solid. This results from the structural nonequivalence of the molecules in the
amorphous solid. Some forces are weaker than others, and when an amorphous material is heated, the weakest intermolecular
9.2.8 https://chem.libretexts.org/@go/page/119768
attractions break first. As the temperature is increased further, the stronger attractions are broken. Thus amorphous materials soften
over a range of temperatures.
Table 9.2.1: Types of Crystalline Solids and Their Properties
Type of Solid Type of Particles Type of Attractions Properties Examples
Figure 9.2.11 : Diamond is extremely hard because of the strong bonding between carbon atoms in all directions. Graphite (in
pencil lead) rubs off onto paper due to the weak attractions between the carbon layers. An image of a graphite surface shows
the distance between the centers of adjacent carbon atoms. (credit left photo: modification of work by Steve Jurvetson; credit
middle photo: modification of work by United States Geological Survey)
You may be less familiar with a recently discovered form of carbon: graphene. Graphene was first isolated in 2004 by using tape
to peel off thinner and thinner layers from graphite. It is essentially a single sheet (one atom thick) of graphite. Graphene,
9.2.9 https://chem.libretexts.org/@go/page/119768
illustrated in Figure 9.2.12, is not only strong and lightweight, but it is also an excellent conductor of electricity and heat. These
properties may prove very useful in a wide range of applications, such as vastly improved computer chips and circuits, better
batteries and solar cells, and stronger and lighter structural materials. The 2010 Nobel Prize in Physics was awarded to Andre
Geim and Konstantin Novoselov for their pioneering work with graphene.
Figure 9.2.12 : Graphene sheets can be formed into buckyballs, nanotubes, and stacked layers.
Crystal Defects
In a crystalline solid, the atoms, ions, or molecules are arranged in a definite repeating pattern, but occasional defects may occur in
the pattern. Several types of defects are known, as illustrated in Figure 9.2.13. Vacancies are defects that occur when positions that
should contain atoms or ions are vacant. Less commonly, some atoms or ions in a crystal may occupy positions, called interstitial
sites, located between the regular positions for atoms. Other distortions are found in impure crystals, as, for example, when the
cations, anions, or molecules of the impurity are too large to fit into the regular positions without distorting the structure. Trace
amounts of impurities are sometimes added to a crystal (a process known as doping) in order to create defects in the structure that
yield desirable changes in its properties. For example, silicon crystals are doped with varying amounts of different elements to
yield suitable electrical properties for their use in the manufacture of semiconductors and computer chips.
Figure 9.2.13 : Types of crystal defects include vacancies, interstitial atoms, and substitutions impurities.
9.2.10 https://chem.libretexts.org/@go/page/119768
Summary
Glossary
amorphous solid
(also, noncrystalline solid) solid in which the particles lack an ordered internal structure
crystalline solid
solid in which the particles are arranged in a definite repeating pattern
interstitial sites
spaces between the regular particle positions in any array of atoms or ions
ionic solid
solid composed of positive and negative ions held together by strong electrostatic attractions
metallic solid
solid composed of metal atoms
molecular solid
solid composed of neutral molecules held together by intermolecular forces of attraction
vacancy
defect that occurs when a position that should contain an atom or ion is vacant
9.2.11 https://chem.libretexts.org/@go/page/119768
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Jim Clark (Chemguide.co.uk)
Daniel James Berger
Wikipedia
Ed Vitz (Kutztown University), John W. Moore (UW-Madison), Justin Shorb (Hope College), Xavier Prat-Resina (University of
Minnesota Rochester), Tim Wendorff, and Adam Hahn.
Adelaide Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
9.2: The Solid State of Matter is shared under a CC BY-SA license and was authored, remixed, and/or curated by LibreTexts.
9.2.12 https://chem.libretexts.org/@go/page/119768
9.2: The Solid State of Matter (Problems)
PROBLEM 9.2.1
At very low temperatures oxygen, O2, freezes and forms a crystalline solid. Which best describes these crystals?
a. ionic
b. covalent network
c. metallic
d. amorphous
e. molecular crystals
Answer
(e) molecular crystals
PROBLEM 9.2.2
As it cools, olive oil slowly solidifies and forms a solid over a range of temperatures. Which best describes the solid?
a. ionic
b. covalent network
c. metallic
d. amorphous
e. molecular crystals
Answer
(d) amorphous
PROBLEM 9.2.3
Explain why ice, which is a crystalline solid, has a melting temperature of 0 °C, whereas butter, which is an amorphous solid,
softens over a range of temperatures.
Answer
Ice has a crystalline structure stabilized by hydrogen bonding. These intermolecular forces are of comparable strength and
thus require the same amount of energy to overcome. As a result, ice melts at a single temperature and not over a range of
temperatures. The various, very large molecules that compose butter experience varied van der Waals attractions of various
strengths that are overcome at various temperatures, and so the melting process occurs over a wide temperature range.
PROBLEM 9.2.4
Identify the type of crystalline solid (metallic, network covalent, ionic, or molecular) formed by each of the following
substances:
a. SiO2
b. KCl
c. Cu
d. CO2
e. C (diamond)
f. BaSO4
g. NH3
h. NH4F
i. C2H5OH
Answer a
9.2.1 https://chem.libretexts.org/@go/page/119769
covalent network
Answer b
ionic
Answer c
metallic
Answer d
molecular
Answer e
covalent network
Answer f
ionic
Answer g
molecular
Answer h
ionic
Answer i
molecular
PROBLEM 9.2.5
Identify the type of crystalline solid (metallic, network covalent, ionic, or molecular) formed by each of the following
substances:
a. CaCl2
b. SiC
c. N2
d. Fe
e. C (graphite)
f. CH3CH2CH2CH3
g. HCl
h. NH4NO3
i. K3PO4
Answer a
ionic
Answer b
Covalent network
Answer c
molecular
Answer d
metallic
Answer e
9.2.2 https://chem.libretexts.org/@go/page/119769
covalent network
Answer f
molecular
Answer g
molecular
Answer h
ionic
Answer i
ionic
PROBLEM 9.2.6
Classify each substance in the table as either a metallic, ionic, molecular, or covalent network solid:
Answer
X = metallic; Y = covalent network; Z = ionic
PROBLEM 9.2.7
Classify each substance in the table as either a metallic, ionic, molecular, or covalent network solid:
Answer
X = ionic; Y = metallic; Z = covalent network
PROBLEM 9.2.8
Substance A is shiny, conducts electricity well, and melts at 975 °C. Substance A is likely a(n):
a. ionic solid
b. metallic solid
c. molecular solid
9.2.3 https://chem.libretexts.org/@go/page/119769
d. covalent network solid
Answer
(b) metallic solid
PROBLEM 9.2.9
Substance B is hard, does not conduct electricity, and melts at 1200 °C. Substance B is likely a(n):
a. ionic solid
b. metallic solid
c. molecular solid
d. covalent network solid
Answer
(d) covalent network solid
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley
(Stephen F. Austin State University) with contributing authors. Textbook content produced by OpenStax College is licensed
under a Creative Commons Attribution License 4.0 license. Download for free at http://cnx.org/contents/85abf193-
[email protected]).
Adelaide Clark, Oregon Institute of Technology
Feedback
Think one of the answers above is wrong? Let us know here.
9.2: The Solid State of Matter (Problems) is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
9.2.4 https://chem.libretexts.org/@go/page/119769
9.3: Structure and General Properties of the Metalloids
Skills to Develop
Describe the general preparation, properties, and uses of the metalloids
Describe the preparation, properties, and compounds of boron and silicon
Explain the properties of some covalent-network solids
Describe semiconductors and some of their properties
A series of six elements called the metalloids separate the metals from the nonmetals in the periodic table. The metalloids are boron, silicon,
germanium, arsenic, antimony, and tellurium. These elements look metallic; however, they do not conduct electricity as well as metals so they are
semiconductors. They are semiconductors because their electrons are more tightly bound to their nuclei than are those of metallic conductors. Their
chemical behavior falls between that of metals and nonmetals. For example, the pure metalloids form covalent crystals like the nonmetals, but like
the metals, they generally do not form monatomic anions. This intermediate behavior is in part due to their intermediate electronegativity values. In
this section, we will briefly discuss the chemical behavior of metalloids and deal with two of these elements—boron and silicon—in more detail.
The metalloid boron exhibits many similarities to its neighbor carbon and its diagonal neighbor silicon. All three elements form covalent
compounds. However, boron has one distinct difference in that its 2s22p1 outer electron structure gives it one less valence electron than it has
valence orbitals. Although boron exhibits an oxidation state of 3+ in most of its stable compounds, this electron deficiency provides boron with the
ability to form other, sometimes fractional, oxidation states, which occur, for example, in the boron hydrides.
Silicon has the valence shell electron configuration 3s23p2, and it commonly forms tetrahedral structures in which it is sp3 hybridized with a formal
oxidation state of 4+. The major differences between the chemistry of carbon and silicon result from the relative strength of the carbon-carbon bond,
carbon’s ability to form stable bonds to itself, and the presence of the empty 3d valence-shell orbitals in silicon. Silicon’s empty d orbitals and
boron’s empty p orbital enable tetrahedral silicon compounds and trigonal planar boron compounds to act as Lewis acids. Carbon, on the other hand,
has no available valence shell orbitals; tetrahedral carbon compounds cannot act as Lewis acids. Germanium is very similar to silicon in its chemical
behavior.
Arsenic and antimony generally form compounds in which an oxidation state of 3+ or 5+ is exhibited; however, arsenic can form arsenides with an
oxidation state of 3−. These elements tarnish only slightly in dry air but readily oxidize when warmed.
Tellurium combines directly with most elements. The most stable tellurium compounds are the tellurides—salts of Te2− formed with active metals
and lanthanides—and compounds with oxygen, fluorine, and chlorine, in which tellurium normally exhibits an oxidation state 2+ or 4+. Although
tellurium(VI) compounds are known (for example, TeF6), there is a marked resistance to oxidation to this maximum group oxidation state.
Figure 9.3.1 : (a) Arsenic and (b) antimony have a layered structure similar to that of (c) graphite, except that the layers are puckered rather than
planar. (d) Elemental tellurium forms spiral chains.
Elemental silicon, germanium, arsenic, antimony, and tellurium are lustrous, metallic-looking solids. Silicon and germanium crystallize with a
diamond structure. Each atom within the crystal has covalent bonds to four neighboring atoms at the corners of a regular tetrahedron. Single crystals
of silicon and germanium are giant, three-dimensional molecules. There are several allotropes of arsenic with the most stable being layer like and
containing puckered sheets of arsenic atoms. Each arsenic atom forms covalent bonds to three other atoms within the sheet. The crystal structure of
antimony is similar to that of arsenic, both shown in Figure 9.3.1. The structures of arsenic and antimony are similar to the structure of graphite,
covered later in this chapter. Tellurium forms crystals that contain infinite spiral chains of tellurium atoms. Each atom in the chain bonds to two
other atoms.
9.3.1 https://chem.libretexts.org/@go/page/119770
Face Centered Cubic Structure (diamond)
Figure 9.3.2 : An icosahedron is a symmetrical, solid shape with 20 faces, each of which is an equilateral triangle. The faces meet at 12 corners.
The name silicon is derived from the Latin word for flint, silex. The metalloid silicon readily forms compounds containing Si-O-Si bonds, which are
of prime importance in the mineral world. This bonding capability is in contrast to the nonmetal carbon, whose ability to form carbon-carbon bonds
gives it prime importance in the plant and animal worlds.
An amorphous substance is a material that appears to be a solid, but does not have a long-range order like a true solid. Treatment with hydrochloric
acid removes the magnesium oxide. Further purification of the boron begins with conversion of the impure boron into boron trichloride. The next
step is to heat a mixture of boron trichloride and hydrogen:
9.3.2 https://chem.libretexts.org/@go/page/119770
1500 °C
Silicon makes up nearly one-fourth of the mass of the earth’s crust—second in abundance only to oxygen. The crust is composed almost entirely of
minerals in which the silicon atoms are at the center of the silicon-oxygen tetrahedron, which connect in a variety of ways to produce, among other
things, chains, layers, and three-dimensional frameworks. These minerals constitute the bulk of most common rocks, soil, and clays. In addition,
materials such as bricks, ceramics, and glasses contain silicon compounds.
It is possible to produce silicon by the high-temperature reduction of silicon dioxide with strong reducing agents, such as carbon and magnesium:
Δ
Extremely pure silicon is necessary for the manufacture of semiconductor electronic devices. This process begins with the conversion of impure
silicon into silicon tetrahalides, or silane (SiH4), followed by decomposition at high temperatures. Zone refining, illustrated in Figure 9.3.3,
completes the purification. In this method, a rod of silicon is heated at one end by a heat source that produces a thin cross-section of molten silicon.
Slowly lowering the rod through the heat source moves the molten zone from one end of the rod to other. As this thin, molten region moves,
impurities in the silicon dissolve in the liquid silicon and move with the molten region. Ultimately, the impurities move to one end of the rod, which
is then cut off.
Silicon reacts with halogens at high temperatures, forming volatile tetrahalides, such as SiF4.
Unlike carbon, silicon does not readily form double or triple bonds. Silicon compounds of the general formula SiX4, where X is a highly
electronegative group, can act as Lewis acids to form six-coordinate silicon. For example, silicon tetrafluoride, SiF4, reacts with sodium fluoride to
yield Na2[SiF6], which contains the octahedral [SiF ] ion in which silicon is sp3d2 hybridized:
6
2−
Antimony reacts readily with stoichiometric amounts of fluorine, chlorine, bromine, or iodine, yielding trihalides or, with excess fluorine or chlorine,
forming the pentahalides SbF5 and SbCl5. Depending on the stoichiometry, it forms antimony(III) sulfide, Sb2S3, or antimony(V) sulfide when
heated with sulfur. As expected, the metallic nature of the element is greater than that of arsenic, which lies immediately above it in group 15.
9.3.3 https://chem.libretexts.org/@go/page/119770
the iodide is a white crystalline solid.
Except for boron trifluoride, the boron trihalides readily hydrolyze in water to form boric acid and the corresponding hydrohalic acid. Boron
trichloride reacts according to the equation:
Boron trifluoride reacts with hydrofluoric acid, to yield a solution of fluoroboric acid, HBF4:
+ −
BF (aq) + HF(aq) + H O(l) ⟶ H O (aq) + BF (aq) (9.3.8)
3 2 3 4
In this reaction, the BF3 molecule acts as the Lewis acid (electron pair acceptor) and accepts a pair of electrons from a fluoride ion:
All the tetrahalides of silicon, SiX4, have been prepared. Silicon tetrachloride can be prepared by direct chlorination at elevated temperatures or by
heating silicon dioxide with chlorine and carbon:
Δ
Silicon tetrachloride is a covalent tetrahedral molecule, which is a nonpolar, low-boiling (57 °C), colorless liquid. It is possible to prepare silicon
tetrafluoride by the reaction of silicon dioxide with hydrofluoric acid:
SiO (s) + 4 HF(g) ⟶ SiF (g) + 2 H O(l) ΔH ° = −191.2 kJ (9.3.10)
2 4 2
Hydrofluoric acid is the only common acid that will react with silicon dioxide or silicates. This reaction occurs because the silicon-fluorine bond is
the only bond that silicon forms that is stronger than the silicon-oxygen bond. For this reason, it is possible to store all common acids, other than
hydrofluoric acid, in glass containers.
Except for silicon tetrafluoride, silicon halides are extremely sensitive to water. Upon exposure to water, SiCl4 reacts rapidly with hydroxide groups,
replacing all four chlorine atoms to produce unstable orthosilicic acid, Si(OH)4 or H4SiO4, which slowly decomposes into SiO2.
Figure 9.3.4 : Laboratory glassware, such as Pyrex and Kimax, is made of borosilicate glass because it does not break when heated. The inclusion
of borates in the glass helps to mediate the effects of thermal expansion and contraction. This reduces the likelihood of thermal shock, which causes
silicate glass to crack upon rapid heating or cooling. (credit: “Tweenk”/Wikimedia Commons)
The boron atom in B(OH)3 is sp2 hybridized and is located at the center of an equilateral triangle with oxygen atoms at the corners. In solid B(OH)3,
hydrogen bonding holds these triangular units together. Boric acid, shown in Figure 9.3.5, is a very weak acid that does not act as a proton donor but
rather as a Lewis acid, accepting an unshared pair of electrons from the Lewis base OH−:
− + −10
B(OH) (aq) + 2 H O(l) ⇌ B(OH) (aq) + H O (aq) Ka = 5.8 × 10 (9.3.12)
3 2 4 3
9.3.4 https://chem.libretexts.org/@go/page/119770
Figure 9.3.5 : Boric acid has a planar structure with three –OH groups spread out equally at 120° angles from each other.
Heating boric acid to 100 °C causes molecules of water to split out between pairs of adjacent –OH groups to form metaboric acid, HBO2. At about
150 °C, additional B-O-B linkages form, connecting the BO3 groups together with shared oxygen atoms to form tetraboric acid, H2B4O7. Complete
water loss, at still higher temperatures, results in boric oxide.
Borates are salts of the oxyacids of boron. Borates result from the reactions of a base with an oxyacid or from the fusion of boric acid or boric oxide
with a metal oxide or hydroxide. Borate anions range from the simple trigonal planar BO ion to complex species containing chains and rings of
3−
3
three- and four-coordinated boron atoms. The structures of the anions found in CaB2O4, K[B5O6(OH)4]⋅2H2O (commonly written KB5O8⋅4H2O)
and Na2[B4O5(OH)4]⋅8H2O (commonly written Na2B4O7⋅10H2O) are shown in Figure 9.3.6. Commercially, the most important borate is borax,
Na2[B4O5(OH)4]⋅8H2O, which is an important component of some laundry detergents. Most of the supply of borax comes directly from dry lakes,
such as Searles Lake in California, or is prepared from kernite, Na2B4O7⋅4H2O.
Figure 9.3.6 : The borate anions are (a) CaB2O4, (b) KB5O8⋅4H2O, and (c) Na2B4O7⋅10H2O. The anion in CaB2O4 is an “infinite” chain.
Silicon dioxide, silica, occurs in both crystalline and amorphous forms. The usual crystalline form of silicon dioxide is quartz, a hard, brittle, clear,
colorless solid. It is useful in many ways—for architectural decorations, semiprecious jewels, and frequency control in radio transmitters. Silica
takes many crystalline forms, or polymorphs, in nature. Trace amounts of Fe3+ in quartz give amethyst its characteristic purple color. The term
quartz is also used for articles such as tubing and lenses that are manufactured from amorphous silica. Opal is a naturally occurring form of
amorphous silica.
The contrast in structure and physical properties between silicon dioxide and carbon dioxide is interesting, as illustrated in Figure 9.3.7. Solid
carbon dioxide (dry ice) contains single CO2 molecules with each of the two oxygen atoms attached to the carbon atom by double bonds. Very weak
intermolecular forces hold the molecules together in the crystal. The volatility of dry ice reflect these weak forces between molecules. In contrast,
silicon dioxide is a covalent network solid. In silicon dioxide, each silicon atom links to four oxygen atoms by single bonds directed toward the
corners of a regular tetrahedron, and SiO4 tetrahedra share oxygen atoms. This arrangement gives a three dimensional, continuous, silicon-oxygen
network. A quartz crystal is a macromolecule of silicon dioxide. The difference between these two compounds is the ability of the group 14 elements
to form strong π bonds. Second-period elements, such as carbon, form very strong π bonds, which is why carbon dioxide forms small molecules with
strong double bonds. Elements below the second period, such as silicon, do not form π bonds as readily as second-period elements, and when they
do form, the π bonds are weaker than those formed by second-period elements. For this reason, silicon dioxide does not contain π bonds but only σ
bonds.
9.3.5 https://chem.libretexts.org/@go/page/119770
Figure 9.3.7 : Because carbon tends to form double and triple bonds and silicon does not, (a) carbon dioxide is a discrete molecule with two C=O
double bonds and (b) silicon dioxide is an infinite network of oxygen atoms bridging between silicon atoms with each silicon atom possessing four
Si-O single bonds. (credit a photo: modification of work by Erica Gerdes; credit b photo: modification of work by Didier Descouens)
At 1600 °C, quartz melts to yield a viscous liquid. When the liquid cools, it does not crystallize readily but usually supercools and forms a glass, also
called silica. The SiO4 tetrahedra in glassy silica have a random arrangement characteristic of supercooled liquids, and the glass has some very
useful properties. Silica is highly transparent to both visible and ultraviolet light. For this reason, it is important in the manufacture of lamps that
give radiation rich in ultraviolet light and in certain optical instruments that operate with ultraviolet light. The coefficient of expansion of silica glass
is very low; therefore, rapid temperature changes do not cause it to fracture. CorningWare and other ceramic cookware contain amorphous silica.
Silicates are salts containing anions composed of silicon and oxygen. In nearly all silicates, sp3-hybridized silicon atoms occur at the centers of
tetrahedra with oxygen at the corners. There is a variation in the silicon-to-oxygen ratio that occurs because silicon-oxygen tetrahedra may exist as
discrete, independent units or may share oxygen atoms at corners in a variety of ways. In addition, the presence of a variety of cations gives rise to
the large number of silicate minerals. Many ceramics are composed of silicates. By including small amounts of other compounds, it is possible to
modify the physical properties of the silicate materials to produce ceramics with useful characteristics.
Video 9.3.2 : A closer look at silicon and its chemistry as a network solid.
Graphite
Carbon has 2 allotropes, or pure elemental forms. The more stable form is graphite, a dark, slippery material used in pencils and lubricants.
Remember that carbon typically makes 4 bonds. The structure of graphite is flat hexagonal sheets; a single sheet is called graphene. Each carbon
atom makes 3 σ bonds and the leftover p orbitals form a delocalized π-bond network over the whole sheet, very similar to the π-bonding in benzene.
The π-bond system actually forms bands, like in a metal, allowing graphite to conduct electricity along the sheets. Weak interactions, like London
dispersion forces, (called π-stacking in this case) hold the sheets loosely together. Because they can slide past each other (especially when impurity
atoms are trapped in between) graphite is a good lubricant.
Diamond Structure
The other bulk allotrope of carbon is diamond. In diamond, each carbon makes 4 bonds in tetrahedral directions to other carbon atoms. The structure
is like the zinc blende ionic structure, except that all the atoms are the same. The properties of diamond (insulator, hard) come from the strong
covalent bonds. Remember that C-C bonds are some of the strongest covalent bonds. It's easiest to think about diamond as forming with sp3 hybrid
9.3.6 https://chem.libretexts.org/@go/page/119770
orbitals on each atom. It's hard to imagine the MO interactions, but the main thing we can know is that because there is good energy match and good
overlap between the atoms, the splitting (energy difference) between bonding and anti-bonding MOs will be big (and there won't be any non-
bonding MOs, because each atom has 4 orbitals and makes 4 bonds). There will still be bands like in metals, but now there are 2 bands with a big
energy gap in the middle. The low-energy, bonding band is called the valence band, and there are exactly enough electrons to fill it. The high-
energy, anti-bonding band is called the conduction band, and it is empty.
If it was hard to understand the previous paragraph, let's just imagine making 1 bond between 2 atoms. We use one sp3 hybrid on each atom. These
point right at each other, have exactly the same energy, and have good overlap. We can imagine drawing an MO diagram just like for H2, with a big
energy gap between the bonding and anti-bonding orbitals. Each atom has 4 electrons total, and makes 4 bonds just like this, so each bond gets 1
electron from each atom: just enough to fill up the bonding MO. When we multiply this over all the bonds in the diamond, we get the full valence
band and the empty conduction band. The energy gap between them is called the band gap. In diamond, the band gap is big, and so diamond is an
insulator.
Video 9.3.3 : A closer look at carbon and its chemistry as a network solid.
Semiconductors
Silicon has a structure just like diamond. (It doesn't make graphite, because mostly π-bonds are weak for the second-row elements.) However, silicon
has a smaller band gap than diamond. Germanium, which is below silicon in the periodic table, has the same structure, and an even lower band gap.
Generally, as we go down the periodic table, covalent bond strengths get smaller, which is the same as saying that the splitting between bonding and
antibonding orbitals gets smaller. It's hard to say why, but if you assume that heavier atoms have worse orbital overlap, you will usually make good
predictions.
Semiconductors are materials that conduct electricity just a little bit. They are the basis for all computing and electronics. Semiconductivity comes
from having a not-quite-full valence band, a not-quite-empty conduction band, or both. If there are a few electrons in the conduction band, they can
conduct electricity just like in a metal (except less, because there aren't very many of them!). If there are a few electrons missing from the valence
band, the empty spots are called holes and they can move around, also conducting electricity. If the band gap of a material is not too big compared to
the thermal energy, then a few electrons can be in the conduction band even though there is room for them in the valence band (because of thermal
energy). For this reason, semiconductors conduct better at higher temperature, because more electrons will be in the conduction band.
9.3.7 https://chem.libretexts.org/@go/page/119770
Figure 9.3.11 : Diagram showing the difference between bands in metals, semiconductors and insulators. EF, the Fermi energy, represents roughly
the cutoff between full and empty MOs: the energy of a state that has a 50% chance of being full at equilibrium. The intrinsic semiconductor has
electrons in the conduction band due to thermal energy.
Doping Semiconductors
The other way to make a semiconductor, with a material whose band gap is too big at room temperature, is by adding some impurity atoms with
different numbers of electrons, which is called doping. Imagine that you have silicon with just a few nitrogen atoms replacing silicon atoms. Each N
has an extra electron. Now there will be a few electrons in the conduction band, so it's a semiconductor! And we can control the conductivity by
controlling how much N we add. We can do the same thing if we add a little bit of boron, which has 3 electrons. In this case, there would be a few
holes in the valence band. These are called p-type (positive, less electrons, like with B) or n-type (negative, extra electrons, like with N)
semiconductors. Lots of important devices, like LEDs, solar cells and transistors (the basis of computer chips) are made of layers of p-type and n-
type semiconductors.
Compound Semiconductors
We can also make semiconductors that are compounds, like gallium arsenide (formula GaAs). These have an AB formula, and there are always 8
valence electrons in the formula. For example, Ga has 3 and As has 5, so GaAs has 8. Another example is ZnSe, (Zn has 2 valence electrons in 4s,
Se has 6 valence electrons just like O). Thus, they have the same number of valence electrons as C or Si. They also have the same structure (zinc
blende, just like diamond except that that atoms alternate types). By combining many different elements, we can change various properties, like band
gap, to be exactly what we want. Generally, the band gap increases as the 2 elements are farther apart (the bonding becomes more ionic, and ionic
solids aren't conductive). We can even mix 3 or more elements, like in CdZnTe (the actual formula would probably be written (Cd,Zn)Te, meaning
that the number of Cd + Zn = Te). To make semiconductors with the right properties, we can use different combinations of elements and also dope
with many different elements.
Summary
The elements boron, silicon, germanium, arsenic, antimony, and tellurium separate the metals from the nonmetals in the periodic table. These
elements, called metalloids or sometimes semimetals, exhibit properties characteristic of both metals and nonmetals. The structures of these
elements are similar in many ways to those of nonmetals, but the elements are electrical semiconductors.
9.3.8 https://chem.libretexts.org/@go/page/119770
Glossary
amorphous
solid material such as a glass that does not have a regular repeating component to its three-dimensional structure; a solid but not a crystal
borate
compound containing boron-oxygen bonds, typically with clusters or chains as a part of the chemical structure
polymorph
variation in crystalline structure that results in different physical properties for the resulting compound
silicate
compound containing silicon-oxygen bonds, with silicate tetrahedra connected in rings, sheets, or three-dimensional networks, depending on the
other elements involved in the formation of the compounds
Contributors
Paul Flowers (University of North Carolina - Pembroke), Klaus Theopold (University of Delaware) and Richard Langley (Stephen F. Austin
State University) with contributing authors. Textbook content produced by OpenStax College is licensed under a Creative Commons Attribution
License 4.0 license. Download for free at http://cnx.org/contents/[email protected]).
Emily V Eames (City College of San Francisco)
Adelaide Clark, Oregon Institute of Technology
Crash Course Chemistry: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
Crash Course Engineering: Crash Course is a division of Complexly and videos are free to stream for educational purposes.
Feedback
Have feedback to give about this text? Click here.
Found a typo and want extra credit? Click here.
9.3: Structure and General Properties of the Metalloids is shared under a CC BY-NC-SA 4.0 license and was authored, remixed, and/or curated by LibreTexts.
9.3.9 https://chem.libretexts.org/@go/page/119770
Index
A E O
aufbau principle electron configuration orbital diagram
3.1: Electron Configurations 3.1: Electron Configurations 3.1: Electron Configurations
C H V
core electron Hund's rule valence electrons
3.1: Electron Configurations 3.1: Electron Configurations 3.1: Electron Configurations
valence shell
3.1: Electron Configurations
1 https://chem.libretexts.org/@go/page/211887
Glossary
Sample Word 1 | Sample Definition 1
1 https://chem.libretexts.org/@go/page/279466
Detailed Licensing
Overview
Title: OIT: CHE 202 - General Chemistry II
Webpages: 72
Applicable Restrictions: Noncommercial
All licenses found:
CC BY-NC-SA 4.0: 51.4% (37 pages)
CC BY 4.0: 36.1% (26 pages)
Undeclared: 8.3% (6 pages)
CC BY-SA 4.0: 4.2% (3 pages)
By Page
OIT: CHE 202 - General Chemistry II - CC BY-NC-SA 4.0 3.3: Trends in Ionization Energy (Problems) - CC
Front Matter - CC BY-NC-SA 4.0 BY-NC-SA 4.0
TitlePage - Undeclared 3.4: Trends in Electron Affinity and Metallic
InfoPage - Undeclared Character - CC BY 4.0
Table of Contents - CC BY-NC-SA 4.0 3.4: Trends in Electron Affinity and Metallic
Licensing - Undeclared Character (Problems) - CC BY-NC-SA 4.0
Unit 1: The Quantum World - CC BY-NC-SA 4.0 Unit 4: Lewis Structures - CC BY-NC-SA 4.0
1.1: The Nature of Light - CC BY 4.0 4.1: Lewis Dot Diagrams - CC BY 4.0
1.1: The Nature of Light (Problems) - CC BY 4.0 4.1: Lewis Dot Diagrams (Problems) - CC BY-
1.2: Atomic Spectroscopy and the deBroglie NC-SA 4.0
Wavelength - CC BY 4.0 4.2: Lewis Structures - CC BY-NC-SA 4.0
1.2: Atomic Spectroscopy and the deBroglie 4.2: Lewis Structures (Problems) - CC BY-NC-SA
Wavelength (Problems) - CC BY 4.0 4.0
Unit 2: Electrons in Atoms - CC BY-NC-SA 4.0 4.3: Formal Charge and Oxidation State - CC BY-NC-
2.1: Coulomb's Law and the Electrostatic Potential - SA 4.0
CC BY 4.0 4.3: Formal Charge and Oxidation State
2.1: Coulomb's Law and the Electrostatic (Problems) - CC BY 4.0
Potential (Problems) - CC BY-NC-SA 4.0 Unit 5: The Strength and Shape of Covalent Bonds - CC
2.2: Atomic Orbitals and Quantum Numbers - CC BY BY-NC-SA 4.0
4.0 5.1: Covalent Bond Formation and Strength - CC BY
2.2: Atomic Orbitals and Quantum Numbers 4.0
(Problems) - CC BY 4.0 5.1: Covalent Bond Formation and Strength
2.3: Atomic Orbitals and the Bohr Model - CC BY 4.0 (Problems) - CC BY-NC-SA 4.0
2.3: Atomic Orbitals and the Bohr Model 5.2: Molecular Shape - CC BY 4.0
(Problems) - CC BY-NC-SA 4.0 5.2: Molecular Shape (Problems) - CC BY-NC-SA
Unit 3: Periodic Patterns - CC BY-NC-SA 4.0 4.0
3.1: Electron Configurations - CC BY 4.0 5.3: Valence Bond Theory and Hybrid Orbitals - CC
3.1: Electron Configurations (Problems) - CC BY- BY 4.0
NC-SA 4.0 5.3: Valence Bond Theory and Hybrid Orbitals
3.2: Trends in Size - CC BY-SA 4.0 (Problems) - CC BY-NC-SA 4.0
Unit 6: Molecular Polarity - CC BY-NC-SA 4.0
3.2: Trends in Size (Problems) - CC BY-SA 4.0
6.1: Electronegativity and Polarity - CC BY 4.0
3.3: Trends in Ionization Energy - CC BY 4.0
1 https://chem.libretexts.org/@go/page/417507
6.1: Electronegativity and Polarity (Problems) - 8.1: Heating Curves and Phase Changes
CC BY-NC-SA 4.0 (Problems) - CC BY-NC-SA 4.0
6.2: Molecular Shape and Polarity - CC BY 4.0 8.2: Solubility and Intermolecular Forces - CC BY 4.0
6.2: Molecular Shape and Polarity (Problems) - 8.2: Solubility and Intermolecular Forces
CC BY-NC-SA 4.0 (Problems) - CC BY-NC-SA 4.0
6.3: Intermolecular Forces - CC BY 4.0 8.3: Concentrations of Solutions - CC BY 4.0
6.3: Intermolecular Forces (Problems) - CC BY- 8.3: Concentrations of Solutions (Problems) - CC
NC-SA 4.0 BY-NC-SA 4.0
Unit 7: Intermolecular and Intramolecular Forces in Unit 9: Semiconductors - CC BY-NC-SA 4.0
Action - CC BY-NC-SA 4.0 9.1: Bond Types and Molecular Orbital Theory - CC
7.1: Surface Tension, Viscosity, and Capillary Action BY 4.0
- CC BY 4.0 9.1: Bond Types and Molecular Orbital Theory
7.1: Surface Tension, Viscosity, and Capillary (Problems) - CC BY-NC-SA 4.0
Action (Problems) - CC BY-NC-SA 4.0 9.2: The Solid State of Matter - CC BY-SA 4.0
7.2: Vapor Pressure - CC BY 4.0 9.2: The Solid State of Matter (Problems) - CC
7.2: Vapor Pressure (Problems) - CC BY-NC-SA BY-NC-SA 4.0
4.0 9.3: Structure and General Properties of the
7.3: Ionic Bond Formation and Strength - CC BY 4.0 Metalloids - CC BY-NC-SA 4.0
7.3: Ionic Bond Formations and Strength Back Matter - CC BY-NC-SA 4.0
(Problems) - CC BY-NC-SA 4.0 Index - Undeclared
Unit 8: Solutions and Phase Changes - CC BY-NC-SA 4.0 Glossary - Undeclared
8.1: Heating Curves and Phase Changes - CC BY 4.0 Detailed Licensing - Undeclared
2 https://chem.libretexts.org/@go/page/417507